Sie sind auf Seite 1von 355

FUNDAMENTAL SCIENCES Chemistry

DYE-SENSITIZED SOLAR CELLS


Edited by K. Kalyanasundaram

With contributions by:


Michael Bertoz, Juan Bisquert, Filippo De Angelis, Hans Desilvestro, Francisco Fabregat-Santiago,
Simona Fantacci, Anders Hagfeldt, Seigo Ito, Ke-jian Jiang, K. Kalyanasundaram, Prashant V. Kamat,
Ladislav Kavan, Jacques-E. Moser, Md. K. Nazeeruddin, Laurence Peter, Henry J. Snaith,
Gavin Tulloch, Sylvia Tulloch, Satoshi Uchida, Shozo Yanagida and Jun-ho Yum

Forewords by:
Michael Grätzel and Shozo Yanagida

EPFL Press
A Swiss academic publisher distributed by CRC Press
CRC Press
Taylor &. Francis Group

Taylor and Francis Group, LLC


6000 Broken Sound Parkway, NW, Suite 300,
Boca Raton, FL 33487
Distribution and Customer Service
orders@crcpress.com

www.crcpress.com

Library of Congress Cataloging-in-PublicationData


A catalog record for this book is available from the Library of Congress.

This book is published under the editorial direction of


Professor Hubert Girault (EPFL).

The publisher, editor and authors of this book would like to thank the Swiss
Federal Institute of Technology (EPFL) for its generous support towards the
publication of this book and are grateful to the following industrial sponsors for
their participation that helped make this project possible:
Dyesol Group, Sefar A.G. and enerStore Consulting, Ltd.

(PflfSHB

is an imprint owned by Presses polytechniques et universitaires romandes, a


Swiss academic publishing company whose main purpose is to publish the
teaching and research works of the Ecole polytechnique fédérale de Lausanne.

Presses polytechniques et universitaires romandes


EPFL - Rolex Learning Center
Post office box 119
CH-1015 Lausanne, Switzerland
E-mail: ppur@epfl.ch
Phone: 021/693 21 30
Fax: 021/693 40 27

www.epflpress.org

© 2010, First edition, EPFL Press, Lausanne (Switzerland)


ISBN 978-2-940222-36-0 (EPFL Press)
ISBN 978-1-4398-0866-5 (CRC Press)
Printed in France
All right reserved (including those of translation into other languages). No part
of this book may be reproduced in any form - by photoprint, microfilm, or any
other means - nor transmitted or translated into a machine language without
written permission from the publisher.
PREFACE
The last century witnessed an incredible number of technological advances that have
changed our lifestyle considerably. The extensive use and growing dependence on
electrical and electronic equipment have increased the energy/power requirements
on a global scale. With dwindling fossil-fuel reserves, there is an urgent need to find
alternative energy resources to meet the growing demand. Alternate energy resources
must be efficient, cost-effective and ecologically friendly. The harnessing of solar
energy, in this context, becomes a very attractive proposition. The sunlight reaching
the earth's surface every day far exceeds the annual demand. A moderately efficient
solar cell array (with 8-10 % efficiency) covering a limited area of the earth's surface
would be able to provide an enormous amount of electric power and thus help reduce
greenhouse-gas emissions.
Chemists have been interested for a long time in the harnessing of sunlight,
either to drive useful chemical transformations or to convert the light directly into
electrical energy. Two short publications in Nature by Honda, Graetzel and cowork-
ers have had a dramatic impact on the focus of research for those chemists interested
in photochemical and conversion and storage of solar energy (A. Fujishima and K.
Honda, Nature, 238, 37 (1972) and B. O'Regan and M. Graetzel, Nature 353, 737
(1991)). The first publication demonstrated the possibility of the photo-decompo­
sition of water into its constituent elements through the irradiation of semiconduc­
tor electrodes such as Ti0 2 immersed in aqueous electrolyte. The second publication
described two important variants of this photo-electrochemical cell, specifically the
use of high-surface-area mesoporous materials for the oxide substrate, and the appli­
cation of dye molecules to harvest the sunlight. Both these propositions have proven
to be seminal to a new field of scientific research.
Interest in the research and development of DSCs is now spread across numer­
ous academic and industrial laboratories. Over six thousand research publications
have appeared in the primary scientific literature on the performance features, and
the number of patents being filed in this area is growing exponentially (already more
than 300 in 2009 for the DSC area alone). The overall solar-to-electrical conversion
efficiency has surpassed 10 % for lab-size cells (under areas of 1 cm2) and 8 % for
modules (25 - 100 cm2). In recognition of the pioneering contributions made by the
Swiss group, the DSC is already referred to as Graetzel Cell. The secondary literature
on DSCs (reviews) is rather limited, most often covering the work of specific research
groups or conference presentations. The DSC is an important contemporary tech­
nology, and one that is rapidly evolving. This monograph presents a comprehensive
introduction to this new emerging area. Indeed, the DSC is the outcome of the cross
XV111 Dye-Sensitized Solar Cells

fertilization of concepts used in photovoltaic solar cells and nanoscience, nanotech-


nology and light-induced electron transfer reactions.
Many features of DSC are unique and advantageous over the solar cells based on
crystalline or amorphous silicon. Nearly all the components of the DSC are "tunable",
including the semiconducting oxide substrate, the dyes, the electrolytes, the redox
mediator and the counter electrode. This has opened great opportunities for chemists
and material scientists. Transparency and multi-color design alone offer huge poten­
tial for the integration of DSCs as part of the building architecture.
The book is organized broadly in two parts. The first half is an overview of the
material choices and performance features of all key components of the DSC. The
second half covers several experimental techniques that help decipher the functioning
of the DSCs in more detail, as well as theoretical calculations that help understand
the key parameters that characterize the performance of the solar cells in quantita­
tive terms. Nearly all the mechanistic studies to quantify parameters that control the
overall performance of the solar cells are discussed. For completeness, the monograph
includes chapters dealing with the scaling-up issues that must be faced to take lab-cell
studies that are academic in nature to the commercialization of the technology in the
form of large-area solar panels and numerous electronic gadgets.
The book benefits from an excellent team of authors, all of whom are experts
with long hands-on experience in various aspects of the DSC technology and have
made seminal contributions to our understanding on how these solar cells operate.
The book is suitable as a text for a one-semester advanced-level course for upper-level
undergraduates and graduate students; it will also serve as a reference work for self-
study for active researchers in the field. In view of the interdisciplinary nature of DSC
science, the book should be of interest to those working in the fields of chemistry,
physics, material science and engineering.
It is a pleasant task to thank all the contributing authors who were kind enough
to spare time from their busy schedule to write the chapters and thus share their exper­
tise with the scientific community at large. At a personal level, it has been a great
privilege for me to be associated in photochemistry research with Prof. Dr. Michael
Graetzel for nearly three decades, sharing both the excitement and agony during the
long period as DSCs matured from one of academic curiosity to an important member
of the family of "third generation solar cells", ready for commercialization in the near
future. Special thanks also go to Dr. Fred Fenter of the EPFL Press for all his help in
putting together this volume.

K. Kalyanasundaram
Lausanne, February 2010
CONTENTS

PREFACE xvii

1 PHOTOCHEMICAL AND PHOTOELECTROCHEMICAL


APPROACHES TO ENERGY CONVERSION 1
K. Kalyanasundaram
1.1 The sun as an abundant energy resource 1
1.2 Photochemical conversion and storage of solar
energy (artificial photosynthesis) 2
1.3 Photographic sensitization 5
1.4 Photoelectrochemical conversion of solar energy 6
1.4.1 Photogalvanic cells 6
1.4.2 Generations of photovoltaic solar cells 7
1.4.3 Photoelectrochemical solar cells with liquid junctions . . .11
1.4.4 Photoredox reactions of colloidal semiconductors
and particulates 14
1.5 Dye sensitization of semiconductors 16
1.5.1 Dye sensitization of bulk semiconductor electrodes 16
1.5.2 Dye-sensitized solar cells - an overview 17
1.5.3 Sequence of electron-transfer steps of a DSC 18
1.5.4 Key efficiency parameters of a DSC 19
1.5.5 Key components of the DSC 21
1.5.6 Quasi-solid state DSCs with spiro-OMeTAD 32
1.5.7 Improvement in efficiency through the
nanostructuring of materials 33
1.5.8 Dye solar cells based on nanorods/nanotubes
and nanowires 34
1.5.9 Sensitization using quantum dots 35
1.5.10 Semiconductor-sensitized ETA solar cells 36
1.5.11 DSCs based on/?-type semiconductor 37
1.6 Conclusions 38
1.7 References 38

2 TITANIA IN DIVERSE FORMS AS SUBSTRATES 45


Ladislav Kavan
2.1 Titania: fundamentals 45
x Dye-Sensitized Solar Cells

2.2 Electrochemistry of titania: depletion regime 48


2.2.1 Photoelectrochemistry under band-gap excitation 49
2.2.2 In-situ FTIR spectroelectrochemistry in the
depletion regime 52
2.2.3 Photoelectrochemistry under sub-band-gap excitation. . . .52
2.3 Electrochemistry of titania: accumulation regime 55
2.3.1 Capacitive processes 56
2.3.2 Li-insertion electrochemistry 57
2.3.3 Spectroelectrochemistry of titania in the accumulation
regime 59
2.4 Titania photoanode for dye sensitized solar cells 60
2.4.1 Non-organized titania made by decomposition of
Ti(IV) alkoxides 61
2.4.2 Electrochemical deposition of titania 62
2.4.3 Aerosol pyrolysis 63
2.4.4 Organized nanocrystalline titania 64
2.4.5 Single-crystal anatase electrode 71
2.4.6 Other methods of producing titania electrodes for DSC . . . 73
2.4.7 Multimodal structures 74
2.5 Conclusion 76
2.6 Acknowledgements 76
2.7 References 76

3 MOLECULAR ENGINEERING OF SENSITIZERS FOR


CONVERSION OF SOLAR ENERGY INTO ELECTRICITY 83
Jun-ho Yum and Md. K. Nazeeruddin
3.1 Introduction 83
3.2 Ruthenium Sensitizers 84
3.2.1 Effect of protons carried by the sensitizers on the
performance 85
3.2.2 Effect of cations in the ruthenium sensitizers on
the performance 86
3.2.3 Device stability 88
3.2.4 Effect of alkyl chains in the sensitizer on
the performance 89
3.2.5 Effect of the ^-conjugation bridge between carboxylic
acid groups and the ruthenium chromophore 92
3.2.6 High Molar Extinction Coefficient Sensitizers 96
3.2.7 Tuning spectral response by thiocyanato ligands 99
3.2.8 Non-thiocyanato ruthenium complexes 101
3.3 Organic sensitizers 102
3.3.1 High efficiency organic sensitizers 102
3.3.2 Near-IR absorbing sensitizers 109
3.4 References 113
Dye-Sensitized Solar Cells xi

4 OPTIMIZATION OF REDOX MEDIATORS AND ELECTROLYTES . .117


Ke-jian Jiang and Shozo Yanagida*
4.1 Introduction 117
4.2 Charge transfer processes in DSCs 118
4.3 Electrolyte components and their roles in the DSCs 121
4.3.1 Organic solvents 121
4.3.2 Cations 121
4.3.3 Additives 123
4.3.4 Electron mediators 125
4.4 Ionic liquid, quasi-solid and solid electrolytes 128
4.4.1 Ionic liquid electrolyte 128
4.4.2 Active iodide molten salts 132
4.4.3 Nonactive iodide molten salts 135
4.4.4 Additives in ILEs 139
4.4.5 Quasi-solid electrolyte 139
4.5 Remarks and prospects 141
4.6 References 142

5 PHOTOSENSITIZATION OF Sn0 2 AND OTHER OXIDES 145


Prashant V. Kamat
5.1 Dependence of the Sensitization Efficiency on the Energy
Difference 146
5.2 Coupled Semiconductor Systems 147
5.3 SnO2-C60-Ru(bpy)f+System 149
5.4 Probing the Interaction of an Excited State Sensitizer
with the Redox Couple 151
5.5 Sensitization of Nanotube Arrays 153
5.6 Charge Separation of Organic Clusters at an Sn0 2 Electrode
Surface 154
5.7 Concluding Remarks 156
5.8 Acknowledgements 156
5.9 References 156

6 SOLID-STATE DYE-SENSITIZED SOLAR CELLS


INCORPORATING MOLECULAR HOLE-TRANSPORTERS 163
Henry J. Snaith
6.1 Introduction 163
6.2 Spiro-OMeTAD-based solid-state dye-sensitized solar cell 165
6.3 The influence of additives upon the solar cell performance 166
6.4 Charge generation: Electron Transfer 168
6.5 Reductive quenching 171
6.6 Charge generation: Hole-transfer 171
6.7 Charge transport in molecular hole-transporters 174
6.8 Hole mobility in spiro-OMeTAD 175
xii Dye-Sensitized Solar Cells

6.9 Influence of charge density on the hole-mobility in molecular


semiconductors 175
6.10 The influence of chemical p-doping upon conductivity and
hole-mobility 177
6.11 The influence of ionic salts on conductivity and hole-mobility . . . 180
6.12 Current collection 181
6.13 Ti0 2 pore filling with molecular hole-transporters 187
6.14 Charge recombination: The influence of additives 192
6.15 Charge recombination: Ion solvation and immobilization 193
6.16 Charge recombination: Controlling the spatial separation
of electrons and holes at the heterojunction 194
6.17 Enhancing light capture in solid-state DSCs 195
6.18 Alternative structures for mesoporous and nanostructured
electrodes in solid-state DSCs 198
6.19 Outlook for hole-transporter based solid-state DSCs 203
6.20 References 203

7 PACKAGING, SCALE-UP AND COMMERCIALIZATION OF


DYE SOLAR CELLS 207
Hans Desilvestro, Michael Bertoz*, Sylvia Tulloch and Gavin Tulloch
7.1 Introduction 207
7.2 From cells to panels 211
7.2.1 Definitions 211
7.2.2 Designs 211
7.2.3 Materials 214
7.2.4 Module performance - experiment vs. modeling 218
7.3 Long-term stability - the key to industrial success 224
7.3.1 Single cells 224
7.3.2 Modules 228
7.3.3 Panels 230
7.4 Scaling up to commercial production levels 231
7.4.1 Material costs and availability 231
7.4.2 Manufacturing 237
7.5 Commercial applications 240
7.6 Conclusions 245
7.7 Acknowledgements 246
7.8 References 246

8 HOW TO MAKE HIGH-EFFICIENCY DYE-SENSITIZED


SOLAR CELLS 251
Seigo Ito
8.1 Introduction 251
8.2 Experimental considerations 252
8.2.1 Preparation of screen-printing pastes 252
8.2.2 Synthesis of Ru-dye 253
Dye-Sensitized Solar Cells xiii

8.2.3 Porous-Ti02 electrodes 254


8.2.4 Counter-Pt electrodes 258
8.2.5 DSC assembling 258
8.2.6 Measurements 260
8.3 Results and discussion 260
8.3.1 TiCl4 treatments 260
8.3.2 Effect of the light-scattering Ti0 2 layer 262
8.3.3 Thickness of the nanocrystalline Ti0 2 layer 263
8.3.4 Anti-reflecting film 263
8.3.5 ReproducibilityofDSCphotovoltaics 264
8.4 Conclusion 265
8.5 Acknowledgements 266
8.6 References 266

9 SCALE-UP AND PRODUCT-DEVELOPMENT STUDIES OF


DYE-SENSITIZED SOLAR CELLS IN ASIA AND EUROPE 267
K. Kalyanasundaram, Seigo Ito, Shozo Yanagida and Satoshi Uchida
9.1 Introduction 267
9.2 Scaling up of laboratory cells to modules and panels 268
9.3 DSC development studies in various European laboratories . . . . 271
9.3.1 Energy Research Centre of the Netherlands (ECN) . . . . 271
9.3.2 Fraunhofer Institute for Solar Energy Systems
(Fraunhofer ISE) 273
9.3.3 G24 Innovation 278
9.3.4 3GSolar, Israel 280
9.4 DSC development studies in various laboratories of Japan 281
9.4.1 Aisin Seiki Co. Ltd. and Toyota Central R&D
Laboratories 281
9.4.2 Fujikura Ltd. (Japan) 287
9.4.3 Peccell Technologies, Inc. (Japan) 290
9.4.4 Sharp Co. Ltd. (Japan) 293
9.4.5 Sony Corporation Ltd. (Japan) 295
9.4.6 Shimane Institute for Industrial Technology (Japan). . . .296
9.4.7 TDK Co., Ltd. (Japan) 297
9.4.8 Eneos Co. Ltd. (Japan) 300
9.4.9 NGK Spark Plug Co., Ltd. (Japan) 301
9.4.10 Panasonic Denko Co. Ltd. (Japan) 303
9.4.11 Taiyo Yuden Co., Ltd. (Japan) 305
9.4.12 Dai Nippon Printing Company 306
9.4.13 Mitsubhishi Paper Mills and Sekisui Jushi
Corporation 306
9.4.14 J-Power Co. Ltd. (Japan) 308
9.5 DSC Development Work in Korea and Taiwan 308
9.5.1 Korean Institute of Science and Technology (KIST). . . .308
9.5.2 Electronics and Telecommunications Research
xiv Dye-Sensitized Solar Cells

Institute(ETRI), Korea 311


9.5.3 Samsung SDI, Korea 311
9.5.4 Industrial Technology Research Institute of
Taiwan (ITRI) 312
9.5.5 J Touch Taiwan 313
9.6 DSC development work in Australia and China 313
9.6.1 Dyesol, Australia 313
9.6.2 Institute of Plasma Physics, Chinese Academy
of Sciences 317
9.7 Conclusion 318
9.8 Acknowledgement 319
9.9 References 319

10 CHARACTERIZATION AND MODELING OF DYE-SENSITIZED


SOLAR CELLS: A TOOLBOX APPROACH 323
Anders Hagfeldt and Laurence Peter
10.1 Introduction 323
10.2 Theoretical background 324
10.2.1 Interfacial electron transfer processes in the DSC 324
10.2.2 Electron trapping in the DSC 328
10.2.3 Electron transport in the DSC 331
10.3 The toolbox 336
10.3.1 Determination of inj ection efficiency and electron
diffusion length under steady-state conditions 336
10.3.2 Electrochemical and spectrolectrochemical
techniques to study the energetics of the
oxide/dye/electrolyte interface 342
10.3.3 Electrochemical measurements with thin layer cells. . . .354
10.3.4 Small-amplitude time-resolved methods 357
10.3.5 Methods based on frequency response analysis 362
10.3.6 Photovoltage decay 374
10.3.7 Determination of density of trapped electrons
inDSCs 376
10.3.8 Measuring the internal electron quasi Fermi level
in the DSC 383
10.3.9 Determining the electron diffusion length using
IMVS and IMPS 386
10.3.10 Photoinduced absorption spectroscopy (PIA) 388
10.3.11 Conclusions 395
10.4 Acknowledgments 396
10.5 Appendix 1 Analytical IMPS solutions 396
10.6 Appendix 2 Numerical solutions of the continuity
equation [10.115] 397
10.7 References 399
Dye-Sensitized Solar Cells xv

11 DYNAMICS OF INTERFACIAL AND SURFACE ELECTRON


TRANSFER PROCESSES 403
Jacques-E. Moser
11.1 Introduction 403
11.2 Energetics of charge transfer reactions 406
11.2.1 Mesoscopic metal oxide semiconductors 406
11.2.2 Dye sensitizer 414
11.3 Kinetics of interfacial electron transfer 416
11.3.1 Charge injection dynamics 416
11.3.2 Charge recombination 430
11.4 Electron transfer dynamics involving the redox mediator 440
11.4.1 Kinetics of interception of dye cations by a
redox mediator 441
11.4.2 Conduction band electron - oxidized mediator
recombination 449
11.4.3 Electron transport in nanocrystalline Ti0 2 films 450
11.5 References 453

12 IMPEDANCE SPECTROSCOPY: A GENERAL INTRODUCTION


AND APPLICATION TO DYE-SENSITIZED SOLAR CELLS 457
Juan Bisquert and Francisco Fabregat-Santiago
12.1 Introduction 457
12.2 A basic solar cell model 461
12.2.1 The ideal diode model 461
12.2.2 Physical origin of the diode equation for a solar cell . . .463
12.3 Introduction to IS methods 466
12.3.1 Steady state and small perturbation quantities 467
12.3.2 The frequency domain 469
12.3.3 Simple equivalent circuits 470
12.4 Basic physical model and parameters of IS in solar cells 480
12.4.1 Simplest impedance model of a solar cell 480
12.4.2 Measurements of electron lifetimes 486
12.5 Basic physical models and parameters of IS in dye-sensitized
solar cells 486
12.5.1 Electronic processes in a DSC 486
12.5.2 The capacitance of electron accumulation in a DSC. . . . 488
12.5.3 Recombination resistance 491
12.5.4 The transport resistance 500
12.6 Transmission line models 508
12.6.1 General structure of transmission lines 508
12.6.2 General diffusion transmission lines 512
12.6.3 Diffusion-recombination transmission line 515
12.6.4 Parameters of the diffusion-recombination model 519
12.6.5 Effect of boundaries on the transmission line 520
xvi Dye-Sensitized Solar Cells

12.7 Applications 523


12.7.1 Liquid electrolyte cells 523
12.7.2 Experimental IS parameters of DSCs 526
12.7.3 Nanotubes 540
12.7.4 Effects of the impedance parameters on the j-V curves . . 543
12.8 Acknowledgments 548
12.9 Appendix: properties of measured DSCs 549
12.10 References 550

13 THEORETICAL AND MODEL SYSTEM CALCULATIONS 555


Filippo De Angelis and Simona Fantacci
13.1 Introduction 555
13.2 Theoretical and computational methods 557
13.2.1 Density Functional Theory (DFT) 557
13.2.2 Basis sets 558
13.2.3 The Car-Parrinello method 558
13.2.4 Solvation effects 558
13.2.5 Excited states 559
13.2.6 Nonadiabatic method 559
13.3 Dye sensitizers 560
13.3.1 Ruthenium(II)-polypyridyl sensitizers 560
13.3.2 Calculations on N3 560
13.3.3 Calculations on other Ru(II)-dye sensitizers 563
13.3.4 Trans-complexes 565
13.3.5 Organic sensitizers 566
13.3.6 Squaraine dyes 567
13.4 Studies of the Ti0 2 substrate 568
13.4.1 Ti0 2 models 568
13.5 Dye sensitizers on Ti0 2 574
13.5.1 Organic dyes on Ti0 2 : adsorption and electron
dynamics 575
13.5.2 Inorganic dyes on Ti0 2 : adsorption and excited
states 579
13.6 Conclusions and perspective 588
13.7 References 589

INDEX 593
CHAPTER 1

PHOTOCHEMICAL AND
PHOTOELECTROCHEMICAL APPROACHES
TO ENERGY CONVERSION
K. Kalyanasundaram

1.1 THE SUN AS AN ABUNDANT ENERGY RESOURCE

A striking feature of contemporary society is the life-style based on machines and gadg­
ets that consume power. Currently the energy-requirement (consumption) estimate for
seven billion people worldwide is about 13 terawatts (TW) and this is expected to go
up by another 10 TW in 40 years time. Available fossil fuel resources are limited and
depleting rapidly. Hence, there is increased global awareness concerning the urgent
need to find alternative energy resources to meet our requirements. Three viable
options are being discussed: carbon-fuel-based sources, nuclear power and renewable
sources, such as solar. The main criticism against carbon-based energy is its impact on
the environment; its use will lead to a substantial increase in atmospheric C0 2 levels,
provoking catastrophic climate changes. On the nuclear front, power needs would
require hundreds of gigawatt (GW)-level nuclear power stations to be built, and yet
no viable method has been found to dispose of the dangerous nuclear fuel wastes. The
third choice, that of renewable energy based on the sun as the source is therefore very
attractive and promising for several reasons. Sunlight is an abundant energy resource
freely available, supplied directly to our home. The amount of solar energy reaching
the surface of the earth is 120,000 TW. Even if a small fraction of sunlight could be
converted to alternative and usable energy forms, there would be no worry about the
energy supply line.
In this monograph, we review in depth one specific emerging technology that
will permit direct conversion of sunlight to electricity. Devices that do this type of
conversion are called photovoltaic solar cells, the term "voltaic" having its origin
in the chemical potential differences that occur in materials following absorption of
sunlight. After a historical review of various approaches that chemists have examined
for the photochemical conversion and storage of solar energy, an overview of the
2 Dye-Sensitized Solar Cells

dye-sensitized solar cell (DSC), the topic of this monograph is presented. The objec­
tive here is to provide an overview of various ancient and contemporary photochemi­
cal energy conversion technologies studied to place the scope of dye-sensitized solar
cells in the right context. Functional features of various key components of the DSC
are elaborated to prepare the setting (background) for more in-depth discussions on
these by leading experts in the following Chapters.

1.2 PHOTOCHEMICAL CONVERSION AND STORAGE OF


SOLAR ENERGY (ARTIFICIAL PHOTOSYNTHESIS)

Scientists have been interested in the idea of harnessing the sun for a long time [1.1-1.7].
They noted several centuries ago that distinct chemical transformations occur when
materials are exposed to the sunlight. Much of this interest has been focused on pho­
tosynthesis, the process by which green plants utilize sunlight to decompose water to
its constituent elements H2 and 0 2 and subsequently use them to convert atmospheric
C0 2 to carbohydrates (sugar):
(1.1)

The process of photosynthesis takes place in the chloroplasts, using chloro­


phyll, the green pigment [1.8]. Extensive research effort has been put in to under­
stand and mimic the primary processes of the photosynthesis mentioned earlier. The
efficiency of conversion of solar radiation to useful biomass (chemical energy) for
photosynthesis is very modest, about 3-6 %. An important aspect of the chloroplast
apparatus is the use of two types of chlorophyll molecules (as antenna and reaction
centers) in an optimized configuration that maximizes the efficiency of the process at
both high and low light conditions.
At the turn of 20th century, Giacomo Ciamician, a chemistry professor at the
University of Bologna, Italy, was fascinated by the ability of plants to harness sun­
light. He was the first scientist to investigate photochemical reactions in a systematic
way. He used the open balcony of the building as a "photoreactor" where hundreds
of bottles and glass pipes containing various substances and mixtures were exposed
to sunrays. At the 1912 meeting of the International Congress of Applied Chemistry,
Ciamician proposed replacing "fossil energy (coal)" with the natural solar radiation
reaching the earth [1.9].
Approaches to "artificial photosynthesis" can be broadly classified to three
types [1.2-1.6]:
(i) homolytic bond fission reactions;
(ii) molecular energy conversion-storage systems; and
(iii) light-induced electron transfer reactions.
The homolytic fission of a chemical bond (reaction 1.2)

(1.2)
Photochemical and photoelectrochemical approaches 3

is a particularly simple photochemical reaction and is always endergonic. The primary


products are highly reactive "free radicals", and they do undergo rapid secondary
reactions utilizing full or part of the energy stored in reaction (1.2). Another difficulty
is that the absorbed photons must have energy greater than the AB bond energy. For
homolytic fission reactions to occur with sunlight photons in the visible (near UV)
wavelength region, the AB bond energy must be less than 300 kJ mole-1, which rules
out any chemical bond. Only a handful of reactions do fit to the above criteria, e.g.,
photolysis of NOC1 to give NO and 1/2 Cl2 and that of FeBr2. The quantum yields of
these reactions are very low and hence of no practical value.
Molecular energy storage reactions are those chemical reactions that lead to
net molecular energy storage. Formation of new bonds, isomerisation or reorganiza­
tion of existing bond framework of the molecule in a unimolecular fashion, or in
bimolecular reactions with potential substrates, are some examples. The reactions
studied so far almost all involve unsaturated organic molecules. The endergonic or
energy-storage feature of these reactions often arises from excessive "strain" induced
in the product molecule, or loss of resonance energy. One system that has been studied
extensively is the photo-isomerisation of derivatives of norbornadiene to quadricy-
clane (reaction 1.3).

(1.3)

The above reaction has been shown to be photosensitized using benzophenone


or copper halides. High energy requirement of such isomerisation reactions demand
photosensitizers that absorb in the UV region. Since the number of such high-energy
photons in the solar radiation is very small, overall solar conversion efficiency of such
systems tend to be very low, typically less than 1 %.
The third type of reactions, referred to as photoinduced electron transfer reac­
tions, involve the transfer of one or more electrons between two reactants following
the absorption of light by one:
(1.4)
From now on we use the terms dye (D) and photosensitizer (S) for those reac­
tants in a multi-component system that absorb the light and initiate the energy-con­
version processes. Either the donor or acceptor can act as the light absorber, or light
absorption is achieved by a third component. Absorption of light by molecules and
complexes S raises them to higher electronically excited state S*, where the light
energy is transformed and stored in the form of enhanced reactivity in the electroni­
cally excited state S*.
Numerous studies have established that molecules in the electronically
excited state S* are distinct, with their own structural and electronic properties and
enhanced reactivity [1.10-1.12]. The molecule in the excited state S* readily undergo
4 Dye-Sensitized Solar Cells

bimolecular electron-transfer reactions with suitable donors or acceptor molecules, as


shown in the reactions given below:
(1.5)
(1.6)
With reference to the sensitizer S that absorbs the light, reactions (1.5) and (1.6)
are labeled as oxidative and reductive quenching of the excited state. Light-induced
electron-transfer reactions are also referred to as photoredox reactions. Photoredox
reactions of this type convert a major part of the light energy of the absorbed photon
into chemical energy stored in the products. The electron-transfer products, rich in
energy, have a tendency to undergo back-electron reactions, resulting in rapid re-
establishment of the ground states of the reactants S and D or A:
-heat (1.7)
■heat (1.8)
Weller, Mataga and coworkers first demonstrated the occurrence of this kind of
excited-state electron transfer in organic molecules in the 1960s. The pioneering work
of Rehm and Weiler established free-energy relationships to quantitatively explain
rates and efficiencies of photo-induced electron-transfer reactions. The Rehm-Weiler
relationship is used widely to explain orders of magnitude variation in the rate con­
stants of electron-transfer processes and yields of redox products in terms of the driv­
ing force of the reactions of interest. The reason quenching rate constants approach
a plateau value at the diffusion controlled limit at large driving force (exothermic) is
explained within Marcus theory as due to behavior in the inverted region. Soon there­
after, Balzani, Sutin, Meyer and others demonstrated similar photoredox processes in
numerous transition metal complexes. The most well studied paradigm is the tris(2,2'-
bipyridine) complex of Ru(II), Ru(bpy)3+ and their variants [1.13].
A long-term objective of research into photochemical redox reactions is to
obtain overall generation of fuels such as H2, CH4, and CH3OH. Table 1.1 lists a

Table 1.1 Some of the chemical reactions with net storage energy [1.2-1.6].
Reaction # Electrons AE (V)

H 2 0(l)^H 2 (g) + i o 2 ( g ) 2 1.23

C0 2 (g)^CO(g) + io 2 (g) 2 1.33


C02(g) + H20(1) -> HCOOH (1) + - 02(g) 2 1.48
C02(g) + H20(1) -> HCHO(g) + 02(g) 4 1.35
C02(g) + 2H20 (1) -> CH3OH(l) + - 02(g) 6 1.21
C02(g) + 2H20(l) -> CH4(g) + 20 2 (g) 8 1.06
N2(g) + 3H20(1) -> 2NH3(g) + 1 0 2 (g) 6 1.17
C02(g) + H20(1) -> 1/6 C6H1206(s) + 0 2 (g) 4 l .24
Photochemical and photoelectrochemical approaches 5

number of target chemical reactions that generate fuels via light-induced electron-
transfer processes. Within the framework of "artificial photosynthesis", a reac­
tion of utmost importance is the decomposition of water to its elements H2 and 0 2
and reduction of C0 2 . Most of the reactants listed are transparent to solar radia­
tion. Hence one needs to use an "external" photosensitizer to achieve the over­
all conversion. A second serious problem in achieving the listed reactions is the
"multi-electron transfer" nature of these processes. Without the use of a suitable
"catalyst" to mediate, the desired products are not obtained in reasonable yield.
Water photolysis to its constituents H2 and 0 2 is the Holy Grail for chemists
working in this field. But there are other solar-chemical conversions that are less chal­
lenging. For example decomposition of HX (X = Br, I) to H2 and X2 using solar radia­
tion would be industrially more beneficial than the decomposition of water.

1.3 PHOTOGRAPHIC SENSITIZATION

Becquerel laid the foundations of the field of photoelectrochemistry way back in 1839
with his observation of measureable current passing between two platinum electrodes
in the presence of sunlight when the electrodes are immersed in an electrolyte contain­
ing metal halide salts [1.14]. Moser reported on photosensitization effects in Silver
halide grains in 1877 [1.15]. Silver halide-based photography since then has evolved
as the biggest application of photosensitization phenomenon with several billion dol­
lars global market (until the recent development of digital photography) [1.16]. A
typical photographic film contains tiny crystals of very slightly soluble silver halide
salts such as silver bromide (AgBr), commonly referred to as "grains." The grains are
suspended in a gelatin matrix and the resulting gelatin dispersion (commonly referred
to as an "emulsion") is melted and applied as a thin coating on a polymer base or, as
in older applications, on a glass plate.
When light or radiation of appropriate wavelength strikes one of the silver hal­
ide crystals, a series of reactions begins that produces a small amount of free silver in
the grain. Initially, a free bromine atom is produced after the bromide ion absorbs the
nhoton impartions 1 Q and 1 10V

AgBr (crystal) + hv (radiation) -^ Ag+ + Br + e~ (1.9)


The silver ion can then combine with the electron to produce a silver atom.
(1.10)
Association within the grains produces aggregated species such as Ag2, Ag2, AgJ,
Agg, Ag^, and Ag°4. The free silver produced in the exposed silver halide grains constitutes
what is referred to as the latent image, which is later amplified by the development process.
The grains containing the free silver in the form of Ag°4 are readily reduced by chemicals
referred to as developers, forming relatively large amounts of free silver; the deposit of free
silver produces a dark area in that section of thefilm.The developer under the same condi­
tions does not significantly affect the unexposed grains.
Once the developed image is obtained, a large amount of unexposed and unde­
veloped silver halide remains in the emulsion. If that silver halide is not removed
6 Dye-Sensitized Solar Cells

before the image is exposed to radiation capable of producing a latent image, the image
will continue to darken. The process of removing the residual silver halide from the
image is calledfixing.The silver halides are only slightly soluble in water; therefore, to
remove the material remaining after development it is necessary to convert it to soluble
complexes which can he removed by washing. Sodium thiosulfate, commonly termed
hypo, has been used for this purpose since 1839. The radiation sensitivity of silver hali­
des ends for all practical purposes at about 525 nm. The sensitivity of the silver halides
may be extended to radiation of longer wavelengths by the addition of dyes or "color
sensitizers." Development of digital cameras at low cost without compromise on the
quality has reduced the practice of classical photography to the professionals.

1.4 PHOTOELECTROCHEMICAL CONVERSION OF SOLAR


ENERGY

Photoelectrochemical solar cells (also called photovoltaic cells) are designed to


convert solar radiation directly to electricity. Once stored, electrical energy, as cur­
rent, can be used for many different electrical appliances, including electrolysers for
oxidation or reduction of chemicals. Photovoltaic cells are the most efficient routes
to solar-energy conversion and storage. Herein we review some of the approaches
that have been examined in the past: Photogalvanic cells using metal electrodes,
photovoltaic solar cells based on semiconductor electrodes (solid-state devices) and
liquid-junction photoelectrochemical cells where the semiconductor electrodes are
immersed in redox electrolytes. Even though practical applications of photogalvanic
and "wet" photoelectrochemical systems have not been realized until this date, they
have helped identify critical factors that are to be controlled for all successful applica­
tions. The best solar-conversion efficiency for single junction (bandgap) and multi-
junction solar cells are obtained in solar cells made up of ultra-pure materials.

1.4.1 Photogalvanic cells


Soon after the recognition that exposure of certain chemicals to sunlight can cause
oxido-reduction in the 19th century, attempts have been made to capture the energy
stored in such electron-transfer processes. A simple approach is to introduce two
metal electrodes in a solution of suitable "dye" with an electron acceptor and try to
capture the electron transfer products induced by light, prior to their recombination:
(1.11)
(1.12)
Generation of potential difference is known as "galvanic effect" and hence the
term photogalvanic cell is used to describe such photo-electrochemical devices [1.17,
1.18]. Figure 1.2 shows schematically the principles of operation of a photogalvanic
cell. The net effect of such mediated reduction would be the driving of the electron
through an external load and hence the overall conversion of light to electricity. As
Photochemical and photoelectrochemical approaches 7

Fig. 1.2 Schematic presentation of a photogalvanic cell.

early as 1940, Rabinowich proposed such photogalvanic cells for solar energy conver­
sion to electricity [1.17]. One extensively studied example of photogalvanic cell is the
photolysis of thionine and similar xanthene dyes in the presence of ferric (Fe3+) ions
in an electrochemical cell using two Pt electrodes.
Calculations by Albery et al. [ 1.18] showed that it is possible to obtain an overall
rate for conversion of light energy to electricity in the 5-9 % range. However in spite
of focused efforts over several decades, the best efficiency obtained was only 0.03 %.
Two key factors are responsible for the observed low efficiency. First, the metal elec­
trodes employed did not have any selectivity for the reaction desired (reduction of D+
at the cathode and not the oxidation of A~ or vice versa for the anode). Indiscriminate
electrode processes at the metal electrodes effectively reduce the number of elec­
trons that can flow over the external circuit. Secondly, only a small fraction of the
dye present in the vicinity of the cathode contributed to the measured photocurrent.
Electron-transfer products formed in the bulk of the solution underwent recombina­
tion before reaching the two electrodes.
There are also additional problems such as self-quenching of the excited state
in concentrated dye solutions, limited solubility, and thermal- and photo-instability
during extended photolysis. Time-resolved studies showed that the rate of back elec­
tron transfer was much faster than the rate in the forward direction. This results in
only a small shift of the equilibrium towards products and hence the amount of scav-
engeable high energy products. To a limited extent, using scavenging reagents it is
possible to intercept back-reactions kinetically. Attempts to improve the selectivity
of the two electrodes and thinner path length cells produced only marginal improve­
ments. Though attempts of solar light conversion to electricity using photogalvanic
cells did not yield meaningful results, studies nevertheless helped identify key factors
that are to be controlled if photoredox reactions of dye molecules are to be viable
light-energy-conversion systems. We will return to this subject below.

1.4.2 Generations of photovoltaic solar cells


The need for power in the outer space to run communication and military satellites
provided NASA and major American industries extensive funding to develop highly
efficient solar-to-electrical conversion devices based on semiconductor electrodes.
8 Dye-Sensitized Solar Cells

Load and available-area constraints demand solar panels for satellites to have the
highest possible conversion efficiency, even if the costs are very high [1.19-1.31].
The understanding of photo-processes involving semiconductors require, in
turn, an understanding of the primary mechanisms of charge-carrier generation and
mobility in these materials. Herein we briefly review the very basic points; the reader
is referred to specialized monographs for an in-depth discussion. In a bulk crystalline
semiconductor, the highest occupied and lowest unoccupied molecular orbitals
(HUMO and LUMO) of constituent atoms or molecules converge into valence and
conduction bands. In the absence of dopants, the energy level (Fermi level) of the
semiconductor lies half-way between the separation gap of the valence and conduction
bands. Doping with electron-donors (n-doping) makes the material electron-rich, and
the Fermi level moves closer to the conduction band. Similarly, doping with electron-
acceptors (p-doping) depletes the number of electrons available and the Fermi level
moves closer to the valence band. Optical excitation of the semiconductor with light
of energy higher than the bandgap separation of the semiconductor leads to generation
of free charge carriers, electrons (e~) and holes (h+). In a sandwich structure composed
of an ft-doped and/?-doped semiconductor, charge separation occurs due to bending of
the bands in the vicinity of the interface (see Fig. 1.3).
With light, additional carriers are created, and the single Fermi level splits into
two quasi-Fermi levels in the n-type or »-type region respectively. These quasi-Fermi

(1)

(2)

(3)

(4)

Fig. 1.3 Band Picture of n- and /?-type semiconductors with the indication of the Fermi level
(Ef) before (scheme 1) and after joining (scheme 2) resulting in ap-n junction.
Photochemical and photoelectrochemical approaches 9

levels are now split; the higher the light intensity the more they split. Close to the elec­
trode both quasi-Fermi levels collapse toward the majority quasi-Fermi level, where
they are connected. This shift of the Fermi levels in the electrodes represents the open
circuit voltage, which can be approximated by the shift of the minority quasi-Fermi
levels. Such separation of the charge carriers permits selective collection at the collec­
tor electrodes and a net conversion of sunlight to electric power.
Based on the nature of the material, maximum conversion efficiency obtain­
able, and the associated cost of photovoltaic power, Martin Green has grouped vari­
ous photovoltaic solar cells in three major categories (see Fig. 1.4). First generation
Photovoltaics use the highest purity materials with least structural defects (such as
single crystals). The highest power-conversion efficiencies obtained to date are in first
generation PVs. Due to high labor costs for the material processing and the significant
energy input required, cost per watt is also the highest. It is very unlikely that these
systems will allow photoelectric power conversion for less than US $l/watt. In addi­
tion to single component Si and layered semiconductors, binary semiconductors of
II-IV and III-V have been examined.
Second generation devices are based on low-energy, intensive preparation
techniques such as vapor deposition and electroplating. Since it is difficult to prepare
systems without defects, maximum power conversion is lower. Nearly all thin film
photovoltaics fall in this category, and the power cost can be less than 1 $US/watt.
Most efficient examples are solar cells made up of multi-crystalline or amorphous

Fig. 1.4 Classification of photovoltaic solar cells into three categories, based on the nature of
the materials used and associated cost of electric power generation.
10 Dye-Sensitized Solar Cells

Si, CdTe and Cd-In-Ga-Se (CIGS). Table 1.5 provides a summary of the state-of-the
art conversion efficiency reported for various semiconductor-based solar cells of the
first and second generations. These materials are applied as a thin film to a supporting
substrate such as glass or ceramics, reducing material mass and therefore costs.
There have been several theoretical calculations on maximum power conversion
obtainable using solar radiation. The most popular calculation is that of Shockley and
Queisser. Considering photovoltaic solar cells as a one-photon giving one-electron
threshold device, these authors estimate 31 % as maximum under 1 sun illumination
and 40.8 % under maximal concentrated solar light (46,200 suns). During the past
decade several approaches have been suggested to cut down the energy losses and
increase the overall conversion.
In one classification, all photovoltaic systems that can potentially give power
conversion efficiency over and above the Shockley and Queisser limit are labeled as
third generation photovoltaics. Advances in our understanding of solid-solid and solid-
liquid interfaces of various kinds permit now usage of wide variety of quasi-crystalline
and even amorphous materials made out of monodispersed colloids, polymers, gels and
electrolytes. Since there is excellent potential for these photovoltaic systems, based on
Table 1.5 Confirmed solar conversion efficiency for various photovoltaic systems (single
junction), measured under AM 1.5 (100 mW/cm2) at 25 °C. [31].

PV material Efficiency Area Voc ^sc FF Lab, year


(%) (cm2) (V) (ma/cm2) (%)
Si (crystalline) 25.0 + 0.5 4.00 0.705 42.7 82.8 Sandia 1999
Si (multicryst.) 20.4 ± 0.4 1.002 0.664 38.0 80.9 NREL 2004
Si (thin film) 16.7 + 0.4 4.017 0.645 33.0 78.2 FhG-ISE 2001
Si (thin film) 10.5 + 0.3 94.0 0.492 29.7 72.1 FhG-ISE 2007
GaAs (cryst.) 26.1+0.8 0.998 1.038 29.7 84.7 FhG-ISE 2007
GaAs (thin film) 26.1+0.8 1.001 1.045 29.5 84.6 FhG-ISE 2008
GaAs (multicryst.) 18.4 + 0.5 4.011 0.994 23.2 79.7 NREL 1995
InP (cryst.) 22.1+0.7 4.02 0.878 29.5 85.4 NREL 1990
CIGS (cell) 19.4 + 0.6 0.994 0.716 33.7 80.3 NREL 2008
CIGS (sub-module) 16.7 + 0.4 16.0 0.661 33.6 75.1 FhG-ISE 2000
CdTe (cell) 16.7 + 0.5 1.032 0.845 26.1 75.5 NREL 2001
Si (amorph.) 9.5 + 0.3 1.070 0.859 17.5 63.0 NREL 2003
Si (nanocryst.) 10.1+0.2 1.199 0.539 24.4 76.6 JQA 1997
Dye-sensitized 10.4 + 0.3 1.004 0.729 22.0 65.2 AIST-Sharp 2005
(sub-module)
Dye-sensitized 8.2 + 0.3 25.45 0.705 19.1 61.1 AIST-Sharp 2007
(sub-module)
Dye-sensitized 8.2 18.50 0.659 19.9 62.9 AIST-Sony 2008
(sub-module)
Org. polymer 5.15 1.021 0.876 9.39 62.5 NREL 2006
Organic 1.1 232.8 29.3 0.072 51.2 NREL 2008
(sub-module)
FhG-ISE: Fraunhofer Institute for Solar Energy Systems in Freiburg, Germany; NREL: National Renew­
able Energy Laboratory, Golden CO, USA; AIST: National Institute of Advanced Industrial Science and
Technology, Tsukuba, Japan; JQA: Japanese Quality Assurances Association.
Photochemical and photoelectrochemical approaches 11

novel materials and nanotechnology, to deliver solar electric power at very low costs
(0.1-0.5 $US/watt), solar cells based on dye-sensitization, polymer organic-bulk hetero-
junctions, and quantum dots are also referred to as third-generation PV systems.

Third generation solar cells for the highest possible conversion


Examination of the loss mechanisms that lead to the Shockley and Queisser limit in
single-junction solar cells leads to the identification of two effects: the non-usage of
the photons of energy below the bandgap, and the thermalization losses that occur
with photons of energy much higher than the bandgap. Based on theoretical analysis,
several approaches have been proposed to overcome these: hot carrier cells, up/down
convertors, multiple excitonic charge-carrier generation and multi-junction (tandem-
type) solar cells.
The hot carrier cell tackles the major PV loss mechanism of thermalisation of
carriers. The underlying concept is to slow the rate of photoexcited carrier cooling,
caused by phonon interaction in the lattice, to allow time for the carriers to be col­
lected whilst they are still "hot", thus enhancing the voltage of a cell. Luminescent
materials are being investigated that either absorb one high-energy photon and emit
more than one low-energy photon just above the band gap of the solar cell {down-
converters)', the other approach is a material that absorbs more than one low-energy
photon below the band gap of the cell and emits one photon just above the band gap
(up-converters). The important property for these devices is high quantum efficiency,
meaning that they must be very radiatively efficient.
Multiple excitonic charge-carrier generation refers to the formation of multi­
ple excitons per absorbed photon, which can happen when the energy of the photon
absorbed is far greater than the semiconductor band gap. This phenomenon does not
readily occur in bulk semiconductors, where the excess energy simply dissipates away
as heat before it can cause additional electron-hole pairs to form. In semi-conducting
quantum dots, the rate of energy dissipation is significantly reduced, and the charge
carriers are confined within a minute volume, thereby increasing their interactions and
enhancing the probability for multiple excitons to form. A quantum yield of 300 per­
cent has been reported, for example, for 2.9-nm-diameter PbSe (lead selenide) quan­
tum dots when the energy of the photon absorbed is four times that of the band gap. But
multiple excitons start to form as soon as the photon energy reaches twice the band gap.
Quantum dots made of lead sulphide (PbS) have also shown the same phenomenon.

1.4.3 Photoelectrochemical solar cells with liquid junctions

Chemists have studied the behavior of semiconductor electrodes immersed in elec­


trolytes containing suitable electron donor or acceptor molecules. As in the case of
the semiconductor, the electrolyte solution has its own chemical potential defined by
the nature and composition of additives. In the situation of physical contact, a space-
charge layer is built up and the Fermi levels of the semiconductor are influenced,
leading to "band bending". In the case of a photoanode, band-bending in the depletion
region drives any electron that is promoted into the conduction band into the interior
of the semiconductor, and holes in the valance band are driven toward the electrolyte,
12 Dye-Sensitized Solar Cells

Fig. 1.6 Principles of operation of liquid-junction solar cells based on n- and/?-type semicon­
ductor electrodes.

where they participate in an oxidation reaction. Electrons through the bulk drive an
external load before they reach the counter electrode or storage electrode, where they
participate in a reduction process. Figure 1.6 shows schematically various electron
transfer steps involved.
Under illumination and open-circuit conditions, a negative potential is created
at a photoanode, and as a result the Fermi level for the photoanode shifts in the nega­
tive direction, thus reducing the band-bending. Under illumination with increasing
intensity, the semiconductor Fermi level shifts continually toward negative potentials
until the band-bending effectively reduces to zero, which corresponds to the flat-band
condition. At this point, a photoanode exhibits its maximum photovoltage, which is
equal to the barrier height.
With the help of a second electrode, it is possible to affect the regeneration of
the mediator oxidized (or reduced) at the illuminated semiconductor. The net effect
of the illumination in the photoelectrochemical cell is conversion of light energy to
electrical energy in two steps, initially as modified chemical potential of the electro­
lyte and then as electric power (electrons driven through the external circuit). Under
such conditions, the redox mediator is recycled and the electrochemical photovoltaic
cell is referred to as regenerative solar cell. These photovoltaic cells are also referred
to as liquid-junction solar cells to differentiate them from pure solid-state devices.
There have been a number of studies on the mechanistic aspects of regenerative solar
cells. Pioneering work in this area has been by Gerischer, Memming, Bard, Wrighton,
Heller, and their coworkers [1.32-1.35].
Early studies of regenerative solar cells were carried out in aqueous media.
A serious problem often faced by those who use semiconductors is the rapid photo-
decomposition of the semiconductor in aqueous electrolytes. Photogenerated holes
are strong oxidants capable of oxidizing the semiconductor itself (photocorrosion)
or oxidizing water, resulting in the formation of a thin oxide-insulating layer at the
electrolyte interface. Both the thermodynamic and kinetic factors involved in stability
of the semiconductor have been investigated. The problem appears more acute with
n-type materials, where the photogenerated holes, which move to the interface, are
Photochemical and photoelectrochemical approaches 13

capable of oxidizing the semiconductor itself. For example, with ZnO in an aqueous
solution at pH 0 the half-reaction (£°D = + 0.9 V vs. NHE)
(1.13)
can occur readily with holes produced at the potential of the valence band edge (-3.0 V
vs. NHE). Thus irradiation of a ZnO electrode in an aqueous solution will cause at
least partial decomposition of the semiconductor electrode.
The II-IV group semiconductors of the chalcogenide family is also susceptible
to serious photocorrosion problems. The decomposition potential for CdX (X = S or
Se) has been determined to be within the bandgap. Thus photocorrosion becomes a
thermodynamically favoured reaction for photogenerated holes (reaction 1.14):
(1.14)
Attempts were made to stabilize the semiconductor photoelectrodes with non-
aqueous solvents in order to dissolve the electrolyte and redox species. Early work in
non-aqueous solvents, such as acetonitrile and methanol have resulted in cells with
low efficiencies for several reasons, including the higher resistivity of the solvent-
electrolyte; the limited solubility of redox species; and the poor bulk and surface
properties of the semiconductor. However, the coating of the semiconductor with a
layer of conducting polymer has been found to be effective.
For the cadmium chalcogenides, photocorrosion may be suppressed by kinetic
competition in the presence of efficient hole scavengers, such as polysulfide and ferro-
cyanide. In the absence of such species, photocorrosion is the only reaction. Formation
of Cd2+ and Se has been reported and verified by various experimental techniques. In
aqueous KOH, Cd(OH)2, Se, Se032~, and Se042~ have to be considered as potential
corrosion products. Table 1.7 gives representative data on the light-conversion effi­
ciency of regenerative solar cells based on single-crystal electrodes.
Polycrystalline materials are important to consider due to their potentially low
fabrication and materials costs. However, trapping or recombination of charge carriers

Table 1.7 Light-conversion efficiency of regenerative solar cells based on semiconductor


electrodes immersed in redox electrolyte solutions.

Semiconductor Aqueous redox electrolyte Solar conv. Stability Ref.


efficiency (%) (C/cm2 )

w-GaAs 1 M K2Se, 0.01 M K2Se2, 12.0 35,000 [1.36]


1MKOH
p-lnP 0.3 M V3+, 0.05 M V2+, 5 M HC1 11.5 27,000 [1.37]
rc-GaAs0.72Po.28 1 M K2Se 11.0 3000 [1.38]
rc-WSe2 1 M KI, 0.01 M KI3 10.2 40,0000 [1.39]
«-CuInSe2 6MI-,0.1MCu 2 + ,0.1MIn 3 + 10.1 15,000 [1.40]
«-CuInSe2 Ι 3 / Γ mixture, Cu+ 9.5 70,000 [1.41]
?z-MoSe2 1 M Kl, 0.01 M KI3 9.4 50,000 [1.42]
rc-CdSe 1 M Na2S2,1 M NaOH 7.2 20,000
n-WS2 1 M NaBr, 0.01 M Br2 6.0
rc-CdSe Fe(CN)64+ 12.4 unstable
14 Dye-Sensitized Solar Cells

occurs readily at defect sites, causing a dramatic decrease in light-energy conversion.


The magnitude of the anodic photocurrent has been found to depend upon a number
of experimental variables; light intensity, electrolyte concentration, surface roughness
and solution pH all influence the rate of charge injection. Best results are obtained
with the use of single crystals. Most often they are even etched to take away few layers
just prior to measurements.
Texas Instruments has developed [1.43] a large-scale solar-energy chemical-
convertor SCC (TISES) for the photo-electrolysis of HBr to Br2 and H2. The project
has an associated cost of US$ 1000 million and uses Si-based solar cells to generate
electricity, which is then used to run a classical electrolyser. Solar-to-chemical effi­
ciencies of 8.6 % have been both predicted and measured for the electrolysis of 48
percent HBr to hydrogen and bromine by a full anode/cathode array. An individual
cathode solar-to-hydrogen efficiency of 9.5 % has been obtained.
Semiconductor surfaces have been modified to protect low-band-gap materials
against photocorrosion [1.44,1.45]. A self-driven photoelectrochemical cell consisting
of Pt-coated p-InP and Mn-oxide-coated n-GaAs has been demonstrated to operate
at 8.2 % maximum efficiency to generate H2 and 0 2 under simulated sunlight [1.46].
More recently, a two-band gap cell in a tandem arrangement has been used to split
water at 12 % efficiency [1.47]. A multi-junction GaAs-Si cell has been recently used
to drive water splitting with over 18 % solar-to-electrical conversion efficiency [1.48].

1.4.4 Photoredox reactions of colloidal semiconductors and


particulates
Since the 1970s, there have been numerous studies of light-induced electron-transfer
processes involving finely divided semiconductor materials such as colloids or partic­
ulates in aqueous media. In contrast to bulk semiconductor-electrode-based systems,
both forms of photo-generated charge carriers (electrons and holes) reach the surface
and are available for suitable electron-transfer processes. Figure 1.8 schematically
illustrates the situation with the colloids and particulate systems. Most often, aerated
(or oxygenated) solutions are employed, where the electrons are rapidly scavenged
by molecular oxygen to form peroxides. Particulate systems have been studied as a
potential low cost and efficient means of degrading toxic industrial wastes. Advances
in colloid and sol-gel chemistry permit facile synthesis of well-defined monodis­
perse colloids of various semiconductors. Their translucent nature beyond the band-

Fig. 1.8 Schematic illustration of possible electron-transfer processes in semiconductor colloid


and particulate systems.
Photochemical and photoelectrochemical approaches 15

gap absorption permit detailed time-resolved studies of excited-state charge transfer


involving valence-band holes or conduction-band electrons.
In these systems involving finely divided semiconductors, such as monodis-
persed colloids and particulate systems, it is known that reducing the size to few
nanometers has important and dramatic consequences. A priori, finely divided semi­
conductors may appear as "potentially very inefficient systems", given the serious
limitations identified in studies with single-crystal electrodes in terms of their trap
and defect sites. In particulate systems, the available surface area increases by several
orders of magnitude, and hence there is a distinct possibility that this huge enhance­
ment of the surface area could be counter-productive.
In reality, it has been found that light-energy conversion processes can indeed
be efficient in these finely divided systems. Indeed, there are many distinct differences
in the details of the evolution of charge-carrier transport. Most of these arise from
the fact that the smaller, finely divided particles are of comparable dimension with
respect to critical parameters that control the efficiency, such as charge-carrier dif­
fusion length. Thus, the serious limitation of recombination of charge carriers in the
bulk of the semiconductor is nearly non-existent in nano-sized systems. It is possible
to scavenge quantitatively all the photogenerated charge carriers.
Titanium dioxide Ti0 2 is a low-cost readily available semiconductor with a
bandgap of 3.2 eV. When the particulates of this semiconductor are illuminated with
light of wavelength less than 385 nm in aerated solutions, hydroxyl radicals (·ΟΗ)
and Superoxide ions (02~) are produced as the net reactants from the photogenerated
holes and electrons respectively. Both these are strong oxidants capable of oxidizing
other organic and inorganic molecules present in the solution. This process of het­
erogeneous catalytic photooxidation has been developed extensively as an industrial
process for the decontamination of toxic wastes and pollutants [1.49-1.59]. This area
is known as heterogeneous photocatalaysis. Systems based on this approach are used
to remove or destroy low-level pollutants in air and water. The oxidation potential of
hydroxyl radical (·ΟΗ) is 2.8 V relative to the normal hydrogen electrode (NHE),
much higher than that of other substances used for water disinfection: ozone (2.07 V),
H 2 0 2 (1.78 V), HOC1 (1.49 V) and chlorine (1.36 V). Employed as a heterogeneous
catalyst, titanium dioxide is readily recoverable for recycling.
Using aqueous colloidal suspensions of semiconductors, Grätzel, Kamat and
coworkers studied mechanistic details of interfacial electron-transfer processes of
semiconductors with various redox reagents present in the solution [1.60-1.65]. With
excitation by a short laser pulse of energy greater than the bandgap energy of the
semiconductor, it is possible to generate charge carriers in the colloidal particles and
follow their dynamics via time-resolved measurements. These studies have also been
extended to dye sensitization of semiconductor colloids. Dynamics of charge injec­
tion from the excited state of various xanthene and porphyrin dyes have been probed.
With dyes that are ionic, electrostatic forces (attraction or repulsion) play important
roles in determining the degree of association of the dye to the semiconductor surface.
Oxide semiconductors are amphoteric (due to the presence of large number of ioniz-
able groups OH~/0~/OH2+) and the overall charge on the particle can be positive or
negative depending on the solution pH with respect to the point of zero charge (zeta
potential). In some aspects, these fundamental studies of electron-transfer processes
16 Dye-Sensitized Solar Cells

involving colloidal semiconductors have laid the foundation for the adoption of these
processes into a viable dye-sensitized solar cell.

1.5 DYE SENSITIZATION OF SEMICONDUCTORS

1.5.1 Dye sensitization of bulk semiconductor electrodes


Sensitization of large band-gap semiconductors is a logical extension of the numerous
studies made earlier on the fundamentals of the photographic process, on photogal-
vanic cells and on light-energy conversion using liquid-junction or regenerative solar
cells. Studies of dye sensitization of bulk semiconductor electrodes in turn have laid
the foundations for the development of dye-sensitized solar cells. Numerous authors
have contributed to our understanding, but seminal contributions have been made in
particular by Genscher, Calvin, Tributsch, Willig, Spitler, Parkinson, and cowork-
ers [1.66-1.73]. In this work, metal-oxide semiconductors such as ZnO, Ti0 2 , Sn0 2 ,
ln 2 0 3 , and SrTi0 3 were sensitized with ruthenium polypyridyl complexes or organic
dyes such as rhodamine B, rose bengal (xanthenes), fluorescein, and alkylthiacarbo-
cyanines. Through extensive studies of the charge-injection processes under different
dye conditions, mechanistic details have been established.
In particular, luminescence studies on dyes adsorbed onto semiconductor elec­
trodes have shown that the excited states could be efficiently quenched on these sur­
faces. Photosensitization in general can occur via transfer of the excitation energy to
a suitable state/energy level of the acceptor or by electron transfer. With semiconduc­
tors, oxidation of the dye takes place through transfer of an electron from a molecule's
excited energy level to the conduction band of the semiconductor. In an electrochemi­
cal cell using semiconductor as bulk electrodes, the excited-state charge injection
manifests itself as photocurrents, measurable quantitatively under anodic polarisation.
Reduction of the excited state of the dye is also known to occur through a valence-band
mechanism, requiring cathodic polarisation of the electrode for detection. This process
has been observed with semiconductors with high hole mobility, such as GaP or SiC.
Tributsch, Genscher and Calvin pioneered the field when they examined photo­
sensitization of ZnO using chlorophyll derivatives as a model system for the primary
process in photosynthesis. As part of his doctoral thesis work, Spitler later studied the
excited-state charge injection of rose bengal onto a single crystal Ti0 2 (rutile) elec­
trode [1.64]. A quantum efficiency of 4 x 10~3 was measured for the electron injec­
tion from the excited rose bengal dye to the conduction band of the semiconductor.
Parkinson has extended these studies to oxide surfaces of low-index faces of the ana-
tase and rutile forms of Ti0 2 by covalently attaching dyes to the surfaces.
A frequent observation made in the study of the dye sensitization of photocurrent
at semiconductor electrodes is that the quantum efficiency for conversion of absorbed
photons to electrons is low, usually on the order of a few percent. In several recent
examinations of this problem, it was concluded that inefficient sensitization could be
attributed to states at the electrode surface that facilitate the return of the transferred
electron from the solid to the oxidized dye layer, thus quenching the production of
Photochemical and photoelectrochemical approaches 17

photocurrent. In some cases these states have been attributed to hydrolyzed surface of
oxide electrodes such as ZnO or Sn0 2 . It is fairly clear from these examples that any
surface layer on a semiconductor electrode could lead to the efficient quenching of the
photocurrent through a back reaction.
Such surface layers are absent at electrodes made of the group-VI dichalco-
genides. The van der Waals (001) surfaces of these layered semiconductors do not
oxidize nor interact strongly with solvents, and therefore they provide an abrupt inter­
face between the electronic states of an adsorbed dye and the energy bands of the
semiconductor. The saturation of the bonding on these surfaces also prevents any
chemical reactions or hydrogen bonding interactions with the surface. The van der
Waals surfaces of the layered semiconductors n-WS2 and n-WSe2 can be sensitized
with high quantum yields with the infrared absorbing thiapentacarbocyanine dye. A
quantum yield of electrons per absorbed photon of 0.6-0.8 has been measured [1.72-
1.73]. The surface-dye concentration dependence of the sensitized photocurrent has
also been studied in the presence and absence of the supersensitizer hydroquinone.
Adsorbed dye aggregates could be identified and selectively photooxidized. This has
yielded sensitized photocurrent densities in excess of 40 μΑ/cm.

1.5.2 Dye-sensitized solar cells - an overview


In this section we present a broad overview of the dye-sensitized solar cell (DSC),
covering the basic principles of operation and the various key components that are
currently being optimized with respect to solar-to-electrical conversion [1.74-1.91].
Figure 1.9 shows schematically the basic architecture. A DSC is basically a thin-layer
solar cell formed by sandwich arrangements of two transparent conducting oxides
(TCO) electrodes. The main highly colored electrode has a few-micron-thick mesopo-
rous Ti0 2 layer coated with a photosensitizer. The counter-electrode is composed of
islands of finely divided Pt deposited onto another TCO. The inter-layer space is filled
with an organic electrolyte containing a redox mediator, usually a mixture of iodine
and iodide in a low viscosity organic solvent such as acetonitrile. Best solar conver­
sion efficiency obtained for this type of DSC is in the range of 11-12 % for laboratory
scale cells (area < 1 cm2) and around 8.5 % for large-area modules (100 cm2).

Fig. 1.9 A schematic representation of a dye-sensitized solar cell (DSC).


18 Dye-Sensitized Solar Cells

Fig. 1.10 A schematic representation of a quasi-solid-state version of the dye-sensitized solar


cell using a hole-transport material co-deposited onto the mesoporous oxide layer.
A quasi-solid state variation of the DSC has also been developed. This uses
an organic hole transport material, such as a triarylamine, to transport charges
between the photoanode and the counter-electrode. Figure 1.10 shows schematically
the organization of various components in the quasi-solid-state version. The hole
transporter can be an inorganic p-type semiconductor such as CuSCN or an organic
donor molecule. They can be considered as intermediate between the DSCs based
on organic electrolytes and polymer organic solar cells using bulk heterojunctions.
To avoid the undesirable situation of the hole transporter reaching the back collector
electrode, quasi-solid-state versions of the DSC employ a thin underlayer, also made
of oxides. The quasi-solid-state version can be easily adapted for portable electron­
ics. For example it is possible to use TCO layers deposited onto flexible organic
polymer substrates in order to manufacture a fully flexible lightweight version of
the DSC.

1.5.3 Sequence of electron-transfer steps of a DSC


Exposure of this solar-cell assembly to visible light leads to a sequence of reactions.
Figure 1.11 shows schematically these processes. We first consider the reactions that
take place at the anode, where the absorption of the light by the dye S leads to forma­
tion of its electronically excited state S*:
(photoexcitation) (1.15)
The molecule in the excited state can decay back to the ground state or undergo
oxidative quenching, injecting electrons into the conduction band of Ti0 2 .
(1.16)
(1.17)
The injected electrons travel through the mesoporous network of particles to
reach the back-collector electrode to pass through the external circuit. The oxidized dye
is reduced rapidly to the ground state by the donor (iodide) present in the electrolyte:
(regeneration of S) (1.18)
Photochemical and photoelectrochemical approaches 19

Fig. 1.11 Schematic drawing of a DSC showing the principles of operation.

In the absence of a redox mediator to intercept and rapidly reduce the oxidized
dye (S+), recombination with the electrons of the titania layer takes place, without any
measurable photocurrent:
► S (recombination) (1.19)
The electrons reaching the counter-electrode through the external circuit reduce
in turn the oxidized iodide (Γ) so that the entire sequence of electron transfer reac­
tions involving the dye and the redox mediator ( I2-I~) is rendered cyclic:

(regeneration of I ) (1.20)
If cited reactions alone take place, the overall effect of irradiation with sunlight
is to drive the electrons through the external circuit, i.e., direct conversion of sunlight
to electricity.

1.5.4 Key efficiency parameters of a DSC


The spectral response of the dye-sensitized solar cell depends on the absorption prop­
erties of the dye. Characterization of the cell depends on a number of experimen­
tally accessible parameters, including the photocurrent and photopotentials measured
under different conditions (open and closed circuit, under monochromatic light or
sunlight illumination): 70C, Voc, /sc and Vsc. The term incident photon-to-electrical-
conversion efficiency (IPCE) is a quantum-yield term for the overall charge-injection
collection process measured using monochromatic light (single wavelength source).
20 Dye-Sensitized Solar Cells

The photocurrent measured under closed circuit /sc is the integrated sum of IPCE
measured over the entire solar spectrum:

(1.21)

Thus IPCE(A) can be expressed as


(1.22)
2
where Λ is the wavelength, /sc the current at short circuit (mA/cm ) and φ is the inci­
dent radiative flux (W/m2).
The overall sunlight-to-electric-power conversion efficiency of a DSC is given
by the following expression:

(1.23)

Maximum power obtainable in a photovoltaic device is the product of two terms


7max and Vmax. The value of 7max gives the maximum photocurrent obtainable at some
"maximum power point". The Fill Factor FF is defined as the ratio of (7maxVmax / AcKc)·
The four values 7SC, Voc, FF and η are the key performance parameter of the solar cell.
The overall efficiency (r7global) of the photovoltaic cell can be calculated from
the integral photocurrent density (/ph), the open-circuit photovoltage (V0CX the fill fac­
tor of the cell (FF) and the intensity of the incident light (7S = 1000 W/m2)
(1.24)

The measured photocurrent will depend on the light intensity; the efficiency
of charge injection in the excited-state quenching process; the degree of recombina­
tion of electrons with the oxidized dye (S+); and the efficiency of charge transport in
the titania films to the counter-electrodes. Maximum photovoltage obtainable in such
sensitized solar cells is the energy gap between the chemical potential level of the
mediating redox electrolyte and the conduction band level of Ti0 2 .
Figure 1.12 shows representative photocurrent-voltage data for a dye-sensitized
solar cell based on the Ru-bpy sensitizer N945 measured under AM 1.5 sunlight illu­
mination (100 mW cm-2) (red line), along with the dependence of the power conver­
sion efficiency for monochromatic light (blue line). The cell active area is 0.158 cm2.
The inset shows a photocurrent action spectrum. For efficient DSCs with overall con­
version efficiency >10 %, at 1 sun (AM 1.5) irradiation, typical photocurrent is of the
order of 15-20 mA/cm2, the photovoltage in the range 650-750 mV, with a fill factor
of 0.65-0.80.
The overall sunlight-to-electric-power conversion efficiency of a DSC can be
expressed as the product of three key terms:
(1.25)
Where 77abs is the efficiency of light absorption by the dye; η^ is the efficiency
of charge injection from the excited state of the dye; and r/coU is the efficiency of
charge collection in the mesoporous oxide layer. An ideal photosensitizer will be the
Photochemical and photoelectrochemical approaches 21

Fig. 1.12 Photocurrent-voltage curve of a solar cell based on the Ru-bpy sensitizer N945
measured under AM 1.5 sunlight illumination (100 mW cm-2) (red line). The cell active
area is 0.158 cm2. Insert in blue shows the monochromatic wavelength dependence of the
photocurrent.

one that absorbs all sunlight in the visible-near IR region with high absorption cross-
section (coefficient).
The efficiency of charge injection T7inj depends on the number of low-lying elec­
tronic excited states below the conduction-band edge of the oxide semiconductor and
the ability of these states to undergo electron transfer with the titania in preference
to other decay channels of the excited state. For efficient electron transfer with the
titania, a good electronic coupling between the electron-acceptor level of titania and
the highest occupied molecular orbitals (HOMO) of the dye is required. As has been
shown for many photoredox reactions of molecules and coordination compounds in
solution, the energetics of the charge-injection step can be determined by consider­
ing the redox potentials of the dye excited state relative to the acceptor (conduction
band) level of the oxide substrate. A moderate driving force of approximately 200 mV
ensures that the excited-state electron transfer occurs rapidly and quantitatively.

1.5.5 Key components of the DSC

Dye-sensitized solar cells have many components that have to be optimized, both
individually and then again as a component of a highly interactive assembly that
includes substrate glass with the transparent conducting oxide TCO layer; mesopo-
rous titania Ti0 2 layer; underlayer(s); dye; electrolyte solvent; redox mediator; and
a counter-electrode. In the case of solar modules, different modes of interconnection
of individual cells must be taken into account to optimize the active area for power
generation. Here we provide a broad overview of the progress made in each of these
key areas to set the stage for more extensive discussions in later Chapters.
22 Dye-Sensitized Solar Cells

Substrates for the DSC


As mentioned earlier, the DSC has a sandwich structure involving two transparent
conducting oxide-glass substrates. The requirements for the TCO substrate are low
sheet resistance (nearly temperature independent to the high temperatures used for
sintering of the Ti0 2 layer, 450-500 °C); and a high transparency to solar radiation
in the visible-IR region. Typical sheet resistance of the TCO used is 5-15 Ω/square.
The cost of TCOs rises steeply with lower sheet resistance and better light transmit-
tance. The cost of these two substrates for the electrodes account for nearly half of the
total cost of the solar cell. Both indium-doped tin oxide (In:Sn02, ITO) and fluorine-
doped tin oxide (F:Sn02, FTO) have been employed. Use of glass substrates confers
good protection against penetration of oxygen or water. But the heavy weight of the
glass renders this form of DSC non-portable, restricting its uses to terrestrial power
generation.
Indium-doped tin-oxide TCO is the most common substrate used in many pho­
tonic and optoelectronic devices. This justifies its mass-production on the industrial
scale. Unfortunately it has been found that the thermal stability of the ITO glass is
not good, with layers peeling off the glass and/or formation of defect sites on the sur­
face that can reduce the solar conversion efficiency of the DSC. Hence, the currently
preferred TCO for DSC application is fluorine-doped tin oxide (F:Sn02). There have
been several studies directed towards improving the efficiency-loss problems related
to the usage of ITO for DSCs.
Only two companies (Pilkington, USA and Asahi, Japan) produce F:Sn0 2
coated glass at the moment. The price largely depends on the total area that is pro­
duced; thus the price is highly uncertain and dependent on the commitment of the two
companies to scale up their production. Today, it is asserted that the price might be
approximately 10 $US/m2, which would correspond to 20 cents/Wp. Since the ohmic
resistance of the TCO is too high (typically 10 Ω/square), additional current collec­
tors like silver fingers have to be applied in modules. These have to be shielded from
the electrolyte by some sealant. Such sealants reduce significantly (by 25 % or more)
the effective area of the cell exposed to sunlight.
In order to obtain optimal adhesive and rheological properties of a mesoporous
film with the TCO substrate, suitable binders are added to the colloidal solutions of
Ti0 2 prior to film deposition by doctor-blade techniques or screen printing. Sintering
of the oxide layer at 450-500 °C gives the film two important properties: the sintering
brings the individual colloidal particles to come into close contact so that the conduct­
ance and charge collection properties of the titania layer are improved; and the aerial
oxidation at elevated temperature removes all the organic matter (potential trap sites)
from the mesoporous film. There have been several studies on the performance of
titania layers deposited using different precursors and coating procedures.
Conducting oxide layers can be deposited on a wide variety of substrates,
including polymer-based plastics. Advantages of such DSCs are low weight, flex­
ibility and preparation protocols that are amenable to established industrial meth­
ods such as roll-to-roll printing. ITO-coated polyethyleneterephalate (ITO-PET) and
polyethylenenaphthalate (ITO-PEN) are well known examples. The disadvantages
of plastic substrates include very limited temperature tolerance (max 150-160 °C),
Photochemical and photoelectrochemical approaches 23

comparatively higher sheet resistance (60 ohm/square for ITO-PET) and permeability
of the plastics to humidity (water and oxygen) over extended outdoor exposure.
A third attractive substrate is thin metal foils, such as titanium or stainless steel.
They have essentially same advantages as the polymer-based and plastic substrates.
With metal substrates, care must be taken against corrosion of the metal by the elec­
trolyte. Also light transmittance will be a serious issue, limiting exposure of the solar
cell to sunlight from only one side.

Ti02 as the photoelectrode


Many wide-bandgap oxide semiconductors (Ti02, ZnO, Sn0 2 , ...) have been exam­
ined as potential electron acceptors for DSCs. Ti0 2 turned out to the most versatile,
delivering the highest solar-conversion efficiency. Ti0 2 is chemically stable, non-
toxic and readily available in vast quantities. It is the basic component of white paints.
Particulate dispersions of Ti0 2 , known as P25, is produced by Degussa by the ton for
the paint industry. Thousands of publications have appeared on the preparation of col­
loidal particles of titania by the sol-gel hydrolysis route using different precursors. An
important requirement for the semiconductor is high transport mobility of the charge
carrier to reduce the electron-transport resistance. ZnO, Sn0 2 , Nb 2 0 5 and many titan-
ates of the general formula MT1O3 have been examined as alternative oxides for the
DSCs. Some of these studies are reviewed in Chapter 5 of this volume.
Ti0 2 has many crystalline forms, with anatase, rutile and brookite being the
easily accessible ones. Rutile has a slightly lower bandgap as compared to anatase and
can absorb a few percent of sunlight in the near-UV region. In the standard version of
DSCs, typical film thickness is 2-15 /mi, and the films are deposited using nanosized
particles of 10-30 nm. The highest solar-conversion efficiency is obtained in double-
layer structures, where an underlay er of thickness 2-4 /mi is first deposited using larger
(200-300 nm) size particles. While the beneficial effects of the underlayer has been
unambiguously established for the quasi-solid-state version of DSC that uses hole
transport materials (HTM), clear trends of beneficial effects have yet to be observed
for the common liquid-electrolyte-based DSCs. A number of experimental approaches
have been used for the deposition of the mesoporous film on the TCO substrate. The
most commonly used are doctor-blading, screen printing and spray-drying.

Procedure for the preparation of DSCs


For the key anode component, a mesoporous film of Ti0 2 (of several micron thick­
ness) is deposited using colloidal particles prepared by the controlled sol-gel hydroly­
sis of Ti-alkoxides. This electrode acts as a sponge and can readily take up a variety
of organic and inorganic dye molecules. The dye is deposited by immersion in a stock
solution for less than an hour. Depending on the method of preparation and material
processing, the oxide layer can be highly translucent (ideal for integration as part of
the building architecture) or opaque. The mesoporous layer consists of well intercon­
nected colloidal particles in the size range of 15-30 nm, and the layer thickness is in
the range of 5 to 15 /mi. Depending on the particle size, the effective surface area (for
dye adsorption), porosity and pore volume (for penetration of the redox electrolytes)
of the oxide layer can vary significantly. For the best photovoltaic performance, post
24 Dye-Sensitized Solar Cells

treatment of the mesoporous oxide layers with TiCl4 is also applied and enhanced
haze obtained by optimization of the light scattering properties of the mesoporous
layer (Chap. 8 of the present volume).

Consequences offractal aspects of mesoporous films


It was mentioned earlier that, in earlier studies of dye sensitization of semiconductor sin­
gle-crystal electrodes, the sunlight-conversion efficiency obtained was quite low (<1 %)
due to two important factors: the excited-state injection process is efficient only at the
monolayer coverage level, and there is poor absorption efficiency (cross section) for vis­
ible light by a monolayer of the dye. An important breakthrough, leading to an enhance­
ment in the conversion efficiency, came about when the available surface area for dye
loading was increased considerably in the mesoporous oxide form of thin films.
Figure 1.13 illustrates this difference by comparing the measured IPCE for sen­
sitization of Ti0 2 in two distinct cases: when Ti0 2 is present as a single crystal and
then as a mesoporous oxide layer. Two distinct improvements can be observed in this
figure. The monochromatic conversion efficiency IPCE increased significantly from
10 to greater than 90 %, due to more efficient light absorption properties of the dye
distributed over the larger area within the mesoporous structure. Secondly, the photo-
current generation is much more efficient in the low-energy region of the dye absorp­
tion - a feature unique to the mesoporous nature of the nanocrystalline layers. Because
the overall photocurrent, 7SC, is given by the area under the IPCE vs. wavelength spec­
trum, the results shown in the figure represent an increase by an order of magnitude.
The term roughness factor or fractal dimension is used to refer to this increased
performance of the mesoporous film; this value can be simply calculated as the ratio
of the measured photocurrent in the two cases cited above, with the assumption that
only those molecules that are in direct contact with the oxide surface are photoac­
tive, and that the remainder merely filter the light. Apart from poor light harvesting,
a compact semiconductor film would need to be n-doped to conduct electrons. In this
case energy-transfer quenching of the excited sensitizer by the electrons in the semi­
conductor would inevitably reduce the photovoltaic conversion efficiency.

Dyes for DSC photosensitizers


Along with the mesoporous oxide layer, a key component of the DSC is the photo-
sensitizer ("dye") that absorbs the solar radiation and injects electrons into the con­
duction band of the oxide substrate. Thousands of small-, medium- and large-sized
molecules have been examined as potential candidates during the past two decades.
In addition to organic molecules, coordination complexes of transition metals with
polypyridine or porphine ligands have been studied.
With reference to the simplified DSC model, it is possible to list a number of
desirable properties for a photosensitizer: strong light absorption in the visible and near-
IR region (for efficient light harvesting); good solubility in organic solvents (for facile
deposition from stock solutions in few hours or less); presence of suitable peripheral
anchoring ligands such as -COOH (to promote the effective interaction of the dye with
the oxide surface and thus the coupling of donor and acceptor levels); suitable disposi­
tion of the HOMO and LUMO of the dye molecule (to permit quantitative injection of
Photochemical and photoelectrochemical approaches 25

Fig. 1.13 A comparison of the monochromatic power conversion efficiency for the same Ru-
dye N719 deposited (a) on the single crystal Ti0 2 (anatase) electrode and (b) on a mesoporous
oxide based DSC.
26 Dye-Sensitized Solar Cells

Fig 1.14 Chemical drawings of "red" and "black" dyes used in DSCs.

Table 1.15 The highest power-conversion efficiencies obtained using different polypyridine
complexes of Ru in standard DSCs using mesoporous Ti0 2 layers and iodide-triiodide redox
electrolyte [1.98].

Dye Surface area (cm2) 7J(%) Voc(V) 7Sc (mA/cm2) FF (%)

N-719 <1 11.2 0.84 17.73 74


N-749 0.219 11.1 0.736 20.9 72
N-749 1.004 10.4 0.72 21.8 65
N-719 1.31 10.1 0.82 17.0 72
N-3 2.36 8.2 0.726 15.8 71
K-19 <1 9.19 0.768 16.4 73
N-945 <1 9.55 0.759 17.25 73
N-621 <1 8.69 0.766 16.22 70
Z-907 <1 7.18 0.70 15.09 68
J-13 <1 7.83 0.703 15.65 71

charges from the electronically excited state); good thermal stability; and good chemi­
cal stability (to retain the chemical identity over repetitive oxido-reduction cycles).
Examples of dyes that function well in DSCs are coordination complexes of
Ru and Os [1.91], squaraines [1.92], porphyrins [1.93], perylenes [1.94], pentacene
[1.95], cyanines [1.96, 1.97], and coumarins [1.98]. The paradigm here is the polypy­
ridine complexes of Ru(II) with one or more carboxylic acid groups as a peripheral
substituent. Because of the complex chemical nomenclature, photosensitizers are
most often referred to by their trivial names (for example, N-3 and N-749 are com­
monly also called "red" and "black" dyes, shown in Fig. 1.14). These red and black
dyes are the benchmark complexes used as standards for the development of other
dyes and components of the DSC. Solar conversion efficiency of over 10 % for lab-
cells have been obtained with many Ru dyes of polypyridine family (see data pre­
sented in Table 1.15 for various Ru-polypyridine complexes developed at the EPFL in
Lausanne). Introduction of long alkyl chains as a substituent in the polypyridine ring
increases the hydrophobicity of the dyes significantly, thereby reducing the stability
Photochemical and photoelectrochemical approaches 27

problems arising from residual water content in the oxide layers. The so-called Z-,
K- and C-series of these dyes are typical examples.
Extension of the 7r-conjugation of polypyridine groups with additional antenna-
type substituent groups leads to substantial increase in the visible-light-absorption
properties and higher molar absorbance. The C101, C103 and Z991 dyes with thi-
ophene-based antenna units illustrate this enhancement process (Fig. 1.16). The
metal-to-ligand charge transfer (MLCT) band in the visible region of these complexes
is red shifted from 530 to 553 nm. The molar extinction coefficient is increased from
14 500 to 21 200 M_1 cnr 1 , allowing the use of thinner titania films. With IPCE
values of >70 % obtainable up to 800 nm, this new generation of dyes permit the
fabrication of DSCs with solar-light-conversion efficiencies well over 10 %. Other
examples of D-π-Α sensitizers are C202, C206, C207, C208 dyes with different tri-
arylamine electron donors in conjugation with a thienothiophene unit as a bridge and
a cyanoacrylic acid acceptor. The J-series of dyes that include triaryl groups, such as
J-13 and J-16, show similar improvements. Many organic dyes with an extended π-
system of the donor-acceptor type also have been found to work efficiently in DSCs.
Encouraging results, with solar-conversion efficiencies in the range of 6-8 %, have
been reported using purely "organic" sensitizers with bis-dimethylfluorenylamino as
the donor and a dithiophene unit as a spacer. Typical examples of this JK-series are
JK2, JK45 and JK46. With indoline dye D205 and C217, solar conversion efficiency
of 9.5 % and 9.8 % respectively have been obtained. Imahori, Bauerle and coworkers
have reviewed some of the exciting developments in this area [1.99a].

Fig. 1.16 Examples of two dyes with 7r-conjugation of polypyridine groups with additional
antenna-type substituent groups.
28 Dye-Sensitized Solar Cells

Fig. 1.17 Dialkyl imidazolium salt used an ionic liquid for non-corrosive DSC electrolyte.
Studies with a number of Ru-complexes of the polypyidine family and metal-
loporphyrins showed the need for suitable anchoring groups, such as carboxyl groups,
to obtain high charge injection efficiencies. In general, dyes with one or more car­
boxyl or phosphonate groups peripheral to the polypyridine ligand give higher charge-
injection efficiency as compared to those without these groups. Dyes with sulphonate
groups do not show such enhancements. At least with polypyridine type complexes,
these groups enable efficient electronic coupling between the donor levels of the dye
and the acceptor levels of the oxide substrate.
For DSC design, it is important to have the option to select from a vast variety
of dyes covering different segments of the solar spectrum, particularly in the vis­
ible-near IR region. Coupled with the option provided by the translucent form of the
mesoporous oxide layer, an entirely new dimension opens for their potential applica­
tion as an integral component of buildings - an area of growing importance in the
photovoltaic market known as building-integrated photovoltaics (BIPV). In Chapter
9 of this book discusses several examples of DSCs as they were developed for use in
indoor electronics and for BIPV applications.

Electrolytes for the DSC


The electrolyte plays a very important role in the DSC by facilitating the transport of
charge between the working and counter electrodes (WE and CE). The ideal electro­
lyte solvent is one that has very low viscosity, negligible vapor pressure, high boil­
ing point and high dielectric properties. From industrial perspective, factors such as
robustness (chemical inertness), environmental sustainability, and easy processing are
also important.
Among the many redox mediators examined, the iodide/triiodide couple has
been found to be the best. Under optimized conditions, the iodide has a concentration
of a few millimolar (and iodine ten times greater), dissolved in a mixture of acetonitrile:
valeronitrile (ACN:VN) or neat N-methoxypriopionitrile MPN. In some studies pro-
pylene carbonate or ethylene carbonate have also been used. Lil had been used as the
iodide source in earlier studies, but now the ionic liquid methylpropylimidazolium
iodide is used. Since the I/I3 mixture also absorbs in the visible-light region, their
concentrations have to be kept as low as possible. The low-viscosity solvent ACN
permits the use of low iodide concentrations.
Large-area solar panels for terrestrial power generation have to meet stringent
stability requirements and must perform reliably. Although ISO or equivalent standards
Photochemical and photoelectrochemical approaches 29

have not yet been proposed, the DSC industry as it is currently under development uses
a set of standards defined for thin-film solar cells, referred to as IEC 61215, IEC 61646,
JIS C-8938 and ASTM E. Included in these standards are a number of thermal- and
light-stability tests: the thermal cycling of the solar cell between 85 and -40 °C to assess
the ability of the panel to withstand exposure to different environmental conditions;
thermal and humidity cycling at 85 % relative humidity at 85 °C; and light soaking
(exposure of the solar cell to continuous illumination under solar light for 1000 hours).
Even though low-viscosity solvents, such as acetonitrile, provide the best solar-
power-conversion efficiency, their low boiling points and high vapor pressures limit
their usage at elevated temperatures (>80 °C), as required by the standards mentioned
above. In a closed environment, the local vapor pressure of the solvent rises to very
high values when their boiling points are reached, and the ability of the solvents to
extract components, even from solids, become very high. In DSCs, this causes slow
extraction of the sealing materials to the electrolyte, accompanied by an associated
decrease in performance of the device. Hence there has been a systematic effort to
find alternatives to low-viscosity, high-vapor-pressure solvents. Simple coordination
complexes of Co, for example, Co(II)/Co(III) have been found to give IPCE values
in the range of 80 % and overall power conversion efficiency of 4 %. Yanagida and
coworkers and others have explored the possibility of using various gelating agents to
create a non-corrosive medium and have obtained very encouraging results [1.99b].
Ionic liquids of the dialkyl imidazolium salts family (also known as room-tem­
perature molten salts, shown in Fig. 1.17) have been found to be excellent substitutes for
this purpose. Ionic liquids have many desirable properties, such as high thermal stability,
negligible vapor pressure, non-flammability and excellent environmental compatibility.
The limitation, however, is their high viscosity, which is one to two orders of magnitude
higher than the volatile solvents. This reduces significantly the diffusion-limited current,
with the associated overall reduction in the power-conversion efficiency. One approach
to overcome this problem is to mix low-viscosity solvents with ionic liquids. Table 1.18
shows the typical composition of two widely used electrolytes for DSC studies.

The counter electrode


To balance the charge and regenerate the key components, the oxidized form of the
mediator needs to be reduced by the electrons flowing through the external circuit

Table 1.18 Typical currently employed solvent-based electrolyte compositions.

Electrolyte type High η Robust

Solvent ACN:VN (3/1) MPN


Iodide compound 1 M PMII 1 M PMII
Iodine 0.03 M I2 0.15 M I2
Additive 1 0.1 M GuSCN 0.1 M GuSCN
Additive 2 0.5 M TBP 0.5 M NBB
Efficiency 105 % (<1 cm2) 9 % (<1 cm2)
ACN = acetonitrile, VN = valeronitrile, MPN = 3-methoxypropionitrile, PMII = l-propyl-3-
methylimidazolium iodide, GuSCN = guanidinium thicocyanate, TBP = tert-butylpyridine, NBB = 1-butyl-
1 //-benzimidazole.
30 Dye-Sensitized Solar Cells

passing through the counter electrode. For the iodide-I2 mixture, the oxidized form
corresponds to triiodide and reduction involves 2 electrons:
(1.26)
To reduce losses, the counter-electrode material should show good electrocata-
lytic properties. The material of choice has been platinum. It suffices to have very
low quantities of platinum deposited onto a TCO (a few /ig/cm2) as finely divided
particles. As an electrode material, Pt shows excellent chemical stability and very low
overpotential for the tri-iodide reduction reaction (1.26).
In early studies, Papageorgiou et al. [1.99] prepared nanosized Pt-metal clusters
by the thermal decomposition of H2PtCl6 from isopropanol on a transparent conduct-
ing-oxide (TCO) coated substrate to maximize the catalytic effect of Pt on the reduc­
tion of iodine or triiodide. Kim et al. [1.100, 1.101] prepared Pt counter electrodes
by direct- and pulse-current electrodeposition, and they confirmed the formation of
uniform Pt nanoclusters of <40 nm composed of 3-nm nanoparticles, when the pulse-
current-electrodeposition method was used, as opposed to the dendritic growth of Pt
that was observed from direct current electrodeposition. Electrodeposited Pt counter
electrodes are potentially applicable for flexible solar cells, since the high-tempera­
ture processing of the Pt CE cannot be used in these cases.
Pt-catalyst T/SP (sold by Solaronix) is a paste containing a chemical platinum
precursor and is intended for screen printing using a polyester mesh. It is first dried
at 100 °C for 10 minutes prior to firing at 400 °C for 30 minutes. After the firing, a
quasi-transparent activated-platinum layer is obtained, suitable for reducing the over-
potentials of the iodide/tri-iodide redox couple. Typically a screen with 90 mesh/cm is
used for obtaining proper catalytic Pt layers. Encouraging results have been obtained
using a new type of counter electrode, prepared by the evaporation of aluminum onto
general glass with the thickness of 700-1100 nm, followed by the sputtering of a plati­
num layer of thickness 300-450 nm using a small ion-sputtering device. Measured
photocurrent was nearly double that of the standard Pt counter electrode. Low sheet
resistance of the CE along with higher reflection properties are cited as possible
reasons for the positive results observed. Kim et al. [1.101] report on the improved
performance and reduced production cost when a novel Pt-NiO counter electrode is
applied using a co-sputtering system; the improvement is due to the increased electro-
catalytic activity and improved adhesion to the substrate.
Kay and Graetzel [1.102] were the first to explore carbon as alternative, less
expensive electroactive material for use as the counter electrode. This work has been
subsequently adopted for carbon black by Murakami et al. [1.103]. Since then, there
have been a number of studies on the use of different forms of carbon as the counter
electrode material for the DSC, including graphite, pyrolytic carbon, carbon black
(CB), and single and multiwalled carbon nanotubes (CNT).
When an activated carbon electrode containing carbon black or poly(3,4-eth-
ylene dioxy-thiophene polystylenesulfonate) (PEDOT) was employed as the counter
electrode of dye-sensitized solar cells, the energy conversion efficiency was improved
compared to that of the solar cells with a sputtered Pt counter electrode. Saito et al. and
Suzuki et al. reported on the application of a chemically produced conducting poly­
mer and carbon nanotubes for use in a counter electrode, respectively [1.104-1.106].
Next Page

Photochemical and photoelectrochemical approaches 31

Lee and coworkers have used nanosize carbon powders deposited onto con­
ducting glass substrates as a DSC counter electrode [1.107]. These authors have also
reported the successful application of multiwall carbon nanotubes (CNT) as electro-
catalysts for triiodide reduction in DSCs [1.108a]. The defect-rich edge planes of
the bamboo-like multiwall CNTs facilitate the electron-transfer kinetics at the coun­
ter electrode-electrolyte interface, resulting in low charge-transfer resistance and
an improved fill factor. In combination with a dye-sensitized Ti0 2 photoanode and
an organic liquid electrolyte, DSCs with a CNT counter-electrode has shown 7.7 %
energy-conversion efficiency under 1 sun illumination (100 mW/cm2, air mass 1.5 G).
The short-term stability test at moderate conditions confirms the robustness of CNT
counter-electrode DSCs.
Conductive plastic substrates, based on polyethylene terephthalate (PET) or
polyethylene naphthalate (PEN) films, have been used as a substrate for the counter-
electrode of DSCs. Recently the CoS-deposited ITO/PEN electrode has been found to
have high catalytic activity for the iodide/triiodide redox couple. Two DSCs based on
Z907 dye and eutectic ionic liquid as the electrolyte, one using a CoS/ITO/PEN film
and the other a Pt/ITO/PEN film as the counter electrode, showed comparable solar-
conversion efficiencies of 6.5 % [1.108b].
An important development in the design of the DSC has been the monolithic
design that attempts to eliminate the need for a second TCO glass substrate. Monolithic
designs are industrially important approaches because they are amenable to semi- or
fully-automatic roll-to-roll printing of solar cells on flexible substrates. Aisen-Seki of
Toyota has been one of the early pioneers in the monolithic design of DSCs. In recent
work, they have prepared monolithic DSCs by applying a porous ZrO layer as a spacer
and insulator between the photoanode and the counter-electrode. Then a thin layer of
catalyst material is introduced that takes part in the redox reaction. Finally, a porous
graphite layer is printed, which serves as the electrically conductive layer and ensures
contact to the second TCO electrode. After sintering of the layers, a metal organic dye is
adsorbed onto the surface of the titanium dioxide. The cell is then filled with electrolyte,
which penetrates into the porous layers of the cell, and encapsulated.
Hinsch and coworkers [1.109a] recently reported on the preparation of three
types of screen-printable catalytic pastes for use as counter electrodes for monolithic
dye solar cells, encapsulated with a glass frit. The electroless bottom-up method or
so-called polyol process has been applied to fabricate thermally stable Sn02:Sb/Pt
and carbon black/Pt nanocomposites. The catalytic and electric properties of these
materials were compared with a new platinum-free type of carbon counter electrode.
The layers containing low platinum amounts (less than 5 mg/cm2) exhibit a very low
charge-transfer resistance of about 0.4 V/cm2. Also, the conductive carbon layer an
shows acceptable charge-transfer resistance of 1.6 V/cm2. Additionally, the catalytic
layer containing porous carbon black reveals excellent sheet resistance below 5 V/
square; this feature has permitted the authors to produce a low-cost counter electrode
that combines suitable catalytic and conductive properties.
Encouraging results on the use of polypyridine complexes of various transition
metals as an alternative to the popular iodide/triiodide redox couple have also been
reported [1.109b-e]. With the combination of N3 dye and [Co(4,4'-di-t-butyl-2,2'-
bpy)3]2+ in γ-butyrolactone, a conversion efficiency of 1.3 % has been obtained. Using
CHAPTER 2

TITANIA IN DIVERSE FORMS AS


SUBSTRATES
Ladislav Kavan

2.1 TITANIA: FUNDAMENTALS

The name 'Titania' denotes several interestingly diverse objects ranging from classi­
cal theatre (Shakespeare's Queen of the faeries in "A Midsummer Night's Dream"),
through astronomy (one of the moons of Uranus) to chemistry (titanium dioxide,
Ti0 2 ) which is the subject of this chapter.
Titanium dioxide has three widely known polymorphs, which also occur in
nature: tetragonal rutile, anatase and orthorhombic brookite (Table 2.1). Rutile is con­
sidered the thermodynamically stable modification of Ti0 2 , as it is approx. 5-12 kJ/
mol more stable than anatase [2.1]. However, this holds only for macroscopic crys­
tals, whereas for nanocrystals smaller than ca. 10-20 nm in size, anatase is the stable
form. (The conclusion regarding the variation in thermodynamic stability of crystals,
depending on their size, is of a general significance; for instance, it is in agreement
with the fact that also nanometer-sized diamonds are more stable than graphite [2.2]).
The conversion of anatase to rutile occurs in the temperature range of ca. 700-1000
°C, depending on the crystal size [2.3] and impurities [2.4]. Brookite obtained its
name after the British mineralogist James Brooke (1771-1857). This substance is
rarely found in nature, in contrast to rutile and anatase. Moreover, the synthesis of
pure brookite is not a simple task [2.5].
The fourth naturally occurring polymorph is monoclinic TI02(B), revealed in
1991 by Banfield et al. [2.6] This mineral is more rare than brookite, and has been
found exclusively in two geological localities near Binntal, Valais, Switzerland. For
the sake of completeness, we should mention that two high-pressure phases, viz.
orthorhombic Ti02(II) (columbite) and monoclinic Ti02(III) (baddeleyite) have also
occasionally been reported to occur in nature. Both these polymorphs have been
disclosed in the Ries Crater, Nördlingen, Bavaria, Germany [2.7, 2.8], and Ti02(II)
was further traced in Chesapeake Bay, Virginia, USA [2.9]. Obviously, the naturally
46 Dye-Sensitized Solar Cells

Table 2.1 Crystalline forms of titania. Representative lattice parameters are quoted, but slightly
different values can also be frequently found depending on the literature source.
Name Space a [nm] b [nm] c [nm] P
group [g/cm3]

rutile P42/mmm 0.4584 — 0.2953 4.13


anatase 14-i/amd 0.3733 - 0.937 3.79
brookite Pbca 0.5436 0.9166 0.5135 3.99
Ti02(B) C2lm 1.2179 0.3741 0.6525 ß= 107.29° 3.64
Ti02(H) - hollandite lAlm 1.0182 - 0.2966 3.46
Ti02(R) - ramsdellite Pbnm 0.4901 0.9453 0.2958 3.87
Ti02(II) - columbite Pbcn 0.461 0.543 0.487 4.33
Ti02(III) - baddeleyite Fljc 0.465 0.493 0.496 ß=992° 4.73
Ti0 2 - fluorite Fwßm 0.4516 - - 4.71
Ti0 2 - pyrite Pa3 0.486 - - 4.65
Ti02(OII)ι - cotunnite Pnma 0.5163 0.2989 0.5966 5.76

occurring high-pressure phases Ti02(II) and Ti02(III) have been grown as a result
of shock-induced phase transformation upon a meteorite impact, indicating that the
shock pressure must have been between 16 and 20 GPa [2.7].
The remaining titania phases (Table 2.1) are synthetic. They are prepared by
two general procedures of which the first involves a chemical conversion from alkali
titanates at ambient conditions (this strategy is called 'chimie douce' which also high­
lights the fact that a great deal of the pioneering work was carried out in France).
The second synthetic procedure is based on a conversion of common forms of Ti0 2
by the action of high pressures, sometimes by a simultaneous action of high tem­
peratures and pressures. In the first case, the topotactic oxidative extraction of alkali
metal from K0.25T1O2 (hollandite) provides the tetragonal Ti02(H) [2.10] (note that the
generic mineral, hollandite, is monoclinic) and, analogously, Li0 5 Ti0 2 (ramsdellite)
is the precursor for orthorhombic Ti02(R) [2.11] (the parent mineral is orthorhom-
bic Mn0 2 ). This preparative procedure is similar to the first synthesis of Ti02(B)
from K2Ti409 reported by Marchand et al. [2.12, 2.13]. Some other layered titan­
ates, viz. Na2Ti307 and Cs2Ti5On, also give Ti02(B) upon ion-exchange and thermal
dehydration [2.14, 2.15]. H2T14O9H2O can be exfoliated into nanosheets, and further
converted into oriented Ti02(B) films [2.16]. Alternatively, it can be transformed to
Ti02(B) by a solvothermal treatment [2.17-2.19]. Ti02(B) has also been observed
during ball milling of anatase [2.20, 2.21]. (Note a certain mismatch in the notation
of Ti02(B) and the Ti02(III) (badelleyite) in Ref. [2.20]). The "β-Ή0 2 " (apparently
Ti02(B)) was observed in titania films fabricated by laser deposition [2.22].
The most common high-pressure phase is Ti02(II), which starts to grow at
5-7 GPa from anatase (the transformation from rutile is also possible, but less efficient
[2.23-2.27]). This phase subsequently converts to Ti02(III) at ca. 15 GPa and remains
stable up to 70 GPa at room temperature. Reports on the other 'post-baddeleyite'
phases are scarce due to inconclusive experimental evidence and theoretical simula­
tions. An exception to this rule is the Ti02(OII) (cotunnite), which grows at pressures
above 60 GPa and temperatures above 1000 K. This titania polymorph has attracted
Titania in diverse forms as substrates 47

considerable attention as a result of it being considered to be the hardest known oxide


[2.28]. The Ti02(II) phase adopts the structure of o-Pb0 2 (columbite) and is therefore
sometimes also labeled o-Ti0 2 . The Ti02(III) takes on the structures of Zr0 2 (bad-
deleyite). The parent crystallographic structures for the other high-pressure titania
phases are: cotunnite (PbCl2), pyrite (FeS2) and fluorite (CaF2) (Table 2.1).
The crystal structures of anatase and rutile are based on a tetragonal symme­
try, in which the Ti4+ ions are 6-fold coordinated to oxygen. The principle difference
between both structures is the arrangement of the oxygen ions. The rutile structure
(space group D4h14- P42/mmm) is based on a slightly distorted hexagonal close pack­
ing of oxygen ions, whereas the oxygen sub-lattice in anatase (space group D4h [2.19]-
IAi/amd) is cubic close-packed. Both structures have edge-sharing Ti0 6 octahedra,
with two and four edges shared in rutile and anatase, respectively. Furthermore, the
octahedra in anatase are slightly more distorted than in rutile. Orthorhombic brookite
(space group D4h19- Pbca) is of a lower symmetry due to its short-range order being
less regular: in brookite none of the six nearest-neighbor Ti-O bond lengths are iden­
tical. The principal structural feature of brookite is a chain of edge-sharing distorted
Ti0 6 octahedra parallel to the c-axis of the orthorhombic lattice, where each octahe­
dron shares three edges.
The crystal structure of Ti02(B) was first determined by Marchand et al. [2.12]
(Note an error in Ref. [2.12], where the Ti02(B) was mismatched with K2Ti8017). The
structure was further refined by theoretical methods [2.29], X-ray [2.15] and neutron
diffraction [2.14]. The monoclinic unit cell (Table 2.1) is isotypic to that of Na x Ti0 2
(bronze), x = 0.2. This is actually the source of the name "Ti02(B)". (Strictly speak­
ing, this name was historically generated in a different way, first for V02(B) [2.30]
and other isomorphic oxides including Ti02(B) [2.31]). The same crystal structure is
also represented by the mineral Na2Fe2Ti6Oi6 (freudenbergite) [2.32]. Hence, Ti02(B)
can equally be called "freudenbergite modification of Ti0 2 ", although this name is
not common. The structure of Ti02(B) is characterized by two edge-sharing Ti0 6
octahedra which are linked to the neighboring doublet of octahedra by corners. This is
demonstrated in Figure 2.2, which also shows characteristic channels running parallel
to the b-axis. These channels give rise to the unique electrochemical behavior of this
titania polymorph (cf. Sections 2.3.1 and 2.4.4).
Ti0 2 is a wide-band-gap semiconductor; the band-gap energy for rutile is 3 eV and
it is 3.2 eV for anatase. The band-gaps of both anatase and rutile are indirect, although
there is also a direct forbidden transition near 3 eV in rutile. The occupied states of
the valence band edge correspond essentially to the O 2p orbitals, and the conduction
band edge arises predominantly from the Ti 3d orbitals. Due to a crystal field of six
oxygen atoms, the 3d states split into t2g and eg components, but the field is not exactly
octahedral as mentioned above. Both natural and synthetic titanium dioxide are usu­
ally slightly oxygen-deficient, Ti02.x (x ~ 0.01). This small oxygen deficiency, which
can also be considered as a Ti3+ impurity, provides n-type doping to the semiconductor
material. Consequently, as an n-type semiconductor, it gives rise to a photoanode in an
electrochemical cell upon excitation across the band gap (cf. Section 2.2).
Data on the electronic structure of brookite are scarcer, and band-gap energies
ranging from 3.1 to 3.4 eV can be found in the literature, without a clear consensus
whether the transition is direct or indirect. Recently, Reyes-Coronado et al. [2.5] reported
48 Dye-Sensitized Solar Cells

Fig. 2.2 A projection of the idealized structure of Ti02(B) perpendicular to the (010) face.
Reprinted with permission from Chem. Mater. 17, 1248 (2005). Copyright 2005 American
Chemical Society.

on an experimental value of 3.13 eV for the indirect band gap of brookite. The elec­
tronic structure of Ti02(B) was calculated by Nuspl et al. [2.33] Analogously to anatase
and rutile, also n-doped Ti02(B) shows photoanodic and photocatalytic [2.18, 2.19]
activity in UV light. The electronic structure of the high-pressure titania polymorphs
(columbite, baddeleyite, fluorite and cotunnite) was calculated by Kuo et al. [2.34] The
fluorite-type titania turned out to have the narrowest band gap, which was attributed to
the coordination number of 8 and the presence of highly symmetrical TiOs polyhedra.
There are numerous practical applications of titania stemming from the proper­
ties of the Ti0 2 surface and/or from the titania nanomaterials of varying morphology.
The properties of titania surfaces have been reviewed in detail by Diebold [2.35].
Titania nanomaterials are the subjects of a review article by Chen and Mao [2.36].
Early work on nanomaterials for light-induced reactions in general and titania for dye
sensitized solar cells (DSCs) in particular were reviewed by Hagfeld and Graetzel
[2.37]. It is not the purpose of this paper to duplicate these reviews, but rather to high­
light the role of titania in state-of-the-art photoanodes for DSC.

2.2 ELECTROCHEMISTRY OF TITANIA: DEPLETION REGIME

Titania and ZnO were the classical materials on which the fundamental issues of semi­
conductor electrochemistry were developed. These research efforts are summarized in
Titania in diverse forms as substrates 49

the reviews by Genscher [2.38], Wrighton [2.39], Bard [2.40] and Finklea [2.41]. For
electrochemical applications, the oxidic semiconductors provide a superior stability
of the electrode/electrolyte interface as compared to that of conventional semiconduc­
tors, such as Si and GaAs. Consequently, the oxides in general and titania in particular
are attractive for electrochemistry and photoelectrochemistry. Although ZnO is a pop­
ular material in the field, Ti0 2 seems to be more favorable for DSC applications. For a
direct comparison, we may recall the paper by Law et al. [2.42], which demonstrated
that only oriented ZnO nanowires could compete with titania. However, the stability
of ZnO in the usual electrolyte solutions for DSC is an issue to be addressed. The
understanding of titania electrochemistry requires a review of charge-transfer reac­
tions under both depletion and accumulation regimes. In the first case, electrons are
depleted from the space charge region and the electronic bands bend upwards. This is
a result of the Fermi level shift upon application of an electrochemical potential that
is positive to the flatband potential. Conversely, electrons accumulate near the surface
if the applied potential is negative to the flatband and the band-bending points down­
wards. The first theme directly adresses the function of dye sensitized solar cells, and
is therefore reviewed in this section.

2.2.1 Photoelectrochemistry under band-gap excitation

The band-gaps of 3.0 eV (rutile) [2.43], 3.2 eV (anatase) [2.43] and (3-3.22) eV
(Ti02(B)) [2.44] translate into the UV-light wavelengths of ca. 390-415 nm, which
can create electron/hole (e7h+) pairs in the material. Upon contact with an electrolyte
solution, a space-charge layer (thickness W) is formed underneath the Ti0 2 surface.
The electric field of the space charge transports e~ to the bulk, and h+ to the surface,
where a photoanodic reaction, such as the oxidation of water, sets in. The incident
photon to current conversion efficiency (IPCE) equals:

(2.1)

where /ph is the photocurrent density, h is Planck's constant, v is the photon frequency,
P is the incident light power density and e is the electron charge. Figure 2.3 and
Table 2.4 show that a value of IPCE as high as 80 % is accessible for anatase, while a
similar performance is traced also for rutile single crystal [2.43, 2.45]. The photocur­
rent density, /ph, attains limiting value at positive potentials, when the band-bending is
large. If this is not the case, one has to consider also the potential (E) dependence:

(2.2)

where a0 is the optical absorption coefficient, ε0 is the permittivity of vacuum, ^ is


the dielectric constant and Nd is the density of donors. This equation is useful for the
determination of ΕΆ in addition to the impedance measurements (Mott-Schottky plot).
Although we trace a small deviation between the impedance and photoelectrochemi-
cal data for an anatase single crystal electrode (Table 2.4), the results confirm that the
£fb is more negative for the (001) face as opposed to the (101) face. If we neglect the
50 Dye-Sensitized Solar Cells

Fig. 2.3 Photocurrent action spectra of single electrodes from anatase (solid points) and rutile
(open points). Electrolyte solution: aqueous IM Na 2 S0 4 (pH 6.4). 0.3 M Lil, pH 2.O. The elec­
trode potential was 1.0 V vs. SCE. Inset: Data replotted for determining the band gap; anatase
solid circles, rutile = open circles. Reprinted with permission from /. Am. Chem. Soc. 118, 6716
(1996). Copyright 1996 American Chemical Society.

Table 2.4 Electrochemical characteristics of anatase. Data for sin£ le crystal electrodes in two
orientations are quoted with the exposed faces (101) and (001). MS = Mott-Schottky plot,
CV = cyclic voltammetry, CHA = chronoamperometry.

Symbol Quantity units (101) (001) Ref.

Em Fiatband potential V,vs. -0.28 -0.34 [2.43, 2.72]


@ pH = 1 (MS) Ag/AgCl
Nd Density of donors (MS) cm -3 2.5-1019 1.2-1019 [2.43, 2.72]
C Interfacial capacitance (CV) ^F/cm2 (5-20) - [2.43]
En Onset of H+ reduction V,vs. -0.22 -0.31 [2.72]
(0.5 M HC1) Ag/AgCl
DLi Diffusion coefficient cm2/s 7-10-1 4 2-10-13 [2.43, 2.72]
for Li+ (CV)
DLi Diffusion coefficient cm2/s 2-10-13 4-10-13 [2.43, 2.72]
for Li+ (CHA)
ΕΆ Activation energy for eV 0.63 0.52 [2.72]
Li+ insertion
k0 Rate constant of cm/s 2-10-9 io-8 [2.43, 2.72]
Li+ insertion (CV)
Eph Onset of UV-photocurrent Y,vs. -0.24 -0.28 [2.43, 2.72]
@pH=l Ag/AgCl
IPCE Photocurrent efficiency % 80 — [2.43]
@ 2=300nm
Titania in diverse forms as substrates 51

e /h+ recombination in the space-charge layer and assume that its thickness W « l/a0
(where IIa0 is the penetration depth of light), then:

(2.3)

Here, A0 is an absorption constant, Lh is the h+-diffusion length and Eg is the


band-gap. By comparing the Eg and flatband potential for rutile and anatase single
crystal electrodes (cf. Table 2.4) [2.43], it is obvious that the increased Eg of anatase
affects only the position of the valence band edge. In other words, the e~/h+pairs can
cause the splitting of water into H2 and 0 2 only in anatase but not in rutile [2.43]. This
conclusion is of fundamental significance in view of the classical paper by Fujishima
and Honda [2.46]. The fact that a complete splitting of water on photoexcited rutile
(without external bias) is unlikely was first articulated by Wrighton [2.39].
Equation (2.3) is useful for determining Eg. The inset in Figure 2.3 shows this
plot for single crystal electrodes, evidencing the expected values of 3.0 eV for rutile
and 3.2 eV for anatase [2.43]. Nanocrystalline materials normally behave like single-
crystal electrodes, but deviations appear for extremely small anatase crystals (1-2 nm),
obtained by electrodeposition (see Section 2.4.2) [2.47] or Ti0 2 nanosheets (Section
2.4.4) [2.48]. Such crystals exhibit an increase (blue shift) of the band gap, AEg, which
matches the general equation for a quantum confinement effect [2.49]:

(2.4)

where μ is the reduced effective mass of the excitons and L corresponds to the crystal
dimension. The subscripts x, y, and z denote the coordinates of the crystal.
A polycrystalline anatase electrode, fabricated for instance by thermal decom­
position of Ti(IV) alkoxides, usually shows a poorer performance than the single
crystals. These electrodes typically exhibit IPCE values of ca. 30 % [2.45]. An excep­
tion to this rule is the Ti0 2 (anatase) obtained by aerosol pyrolysis on an F-doped
Sn0 2 (FTO) support (cf. Section 2.4.3). This electrode is so active with regard to pho-
toanodic water oxidation that one could easily arrive at the physically unrealistic con­
clusion that IPCE > 100 % [2.50]. The paradox is elucidated if we treat the absorption
of UV-light in terms of the interfacial reflectivity, Ry given by the Fresnel equation:

(2.5)

where n,· and η are complex refractive indices of the two materials forming an inter­
face. Assuming the yield of charge injection (photoexcited h+ to the electrolyte solu­
tion) to be 100 %, the IPCE can be approximated as follows:
(2.6)

Here, d is the Ti0 2 him thickness and the indices 1, 2 and 3 denote Ti0 2 , elec­
trolyte solution and FTO, respectively. Qualitatively, this equation demonstrates that
52 Dye-Sensitized Solar Cells

the IPCE is decreased by the reflection loss at the Ti02/electrolyte solution interface
but increased by the reflection at the Ti02/FTO interface. Second, Eq. (2.6) indicates
that the light power penetrating the electrolyte/Ti02 interface is ca. II % larger as
compared to the air/Ti02 interface, which is often (incorrectly) used as a reference for
the light-harvesting efficiency [2.50].

2.2.2 In-situ FTIR spectroelectrochemistry in the depletion regime


In-situ FTIR spectroelectrochemistry has also been used for investigating the pho-
toanodic breakdown processes at the Ti0 2 electrode when illuminated by UV-light
[2.51-2.53]. The photoanodic oxidation of propylene carbonate [2.51], acetonitrile
[2.51], dimethylsulfoxide [2.52] and 3-methyl-oxazolidin-2-one [2.53] was studied
in detail on spray-pyrolyzed Ti0 2 (Section 2.4.3). In all cases, C0 2 was the final
product, but additional intermediates have also been monitored, allowing a discussion
of the breakdown mechanism of these common solvents for aprotic electrolyte solu­
tions [2.51-2.53]. As expected, the photoanodic reactions give different products than
the cathodic ones. For instance, propylene carbonate decomposes photoanodically to
C0 2 , acetone and unidentified polymers, but not to C0 3 nor to propylene (cf. Section
2.3.3.; Eq. 2.19) [2.51, 2.54].

2.2.3 Photoelectrochemistry under sub-band-gap excitation

As shown in Section 2.2.1, titanium dioxide absorbs UV photons with hv > Eg, which,
however, represent only ca. 4 % of the solar power impinging on Earth. Hence, for
photoelectrochemical solar cells, Ti0 2 needs to be sensitized to longer wavelengths,
which constitutes the basis for the discovery of dye-sensitized solar cells. Table 2.5
compiles a historical overview of the development of titania electrodes for DSC. It is
obvious that the history of DSC outlines the optimization of titania substrates as one
of the crucial parameters in enhancement of the solar conversion efficiency.
The spectral sensitization of Ti0 2 was discovered in 1977 by Clark and Sutin
[2.55] and further developed by Goodenough et al. [2.56] These researchers have
demonstrated that a flat Ti0 2 surface of a rutile single crystal can be sensitized by
Ru(bpy)3+ (bpy = 2,2/-bipyridine) in the aqueous electrolyte solution. The crucial
effect consisted in an electron injection from a photoexcited complex, Ru*(bpy)3+,
formed at λα^ = 452 nm, into the conduction band of Ti0 2 . The produced Ru(bpy)3+
is scavenged by a hole acceptor in solution, such as hydroquinone, HQ, to the starting
Ru(bpy)?+, while quinone, Q, is formed [2.57]:
(2.7a)

(2.7b)
The regenerative loop is completed by a reduction of Q to HQ by photogen-
erated electrons [ecß(Ti02)] transmitted to the Pt counter-electrode. The back electron
transfer:
(2.7c)
Titania in diverse forms as substrates 53

Table 2.5 The history of the development of Ti02 photoanodes for dye-sensitized solar cells.
d = titania film thickness, RF = roughness factor.
Year Titania form d [/mi] RF Ref. Note

1977 single crystal - ~1 [2.55] Ru(bpy)32+ in


rutile aqueous solution

1979 single crystal ~1 [2.58] Surface derivatization. For
(rutile?) anatase see Ref. [2.43]
1985 fractal/Ti 10 100 [2.57]
1988 fractal/Ti 20 181 [2.113]
1990 fractal/FTO - - [2.115]
1990 colloidal 2.7 50 [2.109] 8-nm colloidal
particles (non autoclaved)
1991 colloidal 10 780 [2.59] 15-nm particles +
electrodeposition
1993 colloidal 12 1000 [2.163] 15-nm particles and
TiCl4 treatment
1998 underlayer + 4.2 — [2.60] Compact underlayer by
colloidal spray pyrolysis
2003 'double-layer' 10 + 4 [2.173] 20-mn & 400 nm
particles
2005 Pluronic templated 1 466 [2.134] 3-layers film
— —
2005 nanospheres/ [2.171] Lift-off
nanorods transfer technique

2007 nanotube 5.7 [2.159] Anodic titania
(anatase) on Ti
2008 colloidal + -1 + - [2.143] Colloidal particles
inv. opal (3 + 2) overlaid with inverse opal

2008 Pluronic + P25 0.3 + 11 [2.182] Mesoporous underlayer +
nanoparticles (P25)-film

is prevented by band-bending. The sensitization by Ru(bpy)32+ (Eqs. 2.7a,b) is not


very efficient (cf. Table 2.6), and the reason for this is that, for the lifetime of the
Ru*(bpy)32+ (t= 1.74/is) and its diffusion coefficient, D ~ 10~5 cm2/s, the effective
diffusion distance of Ru*(bpy)32+ is [2.45, 2.57]:
(2f-D) 1 / 2 -59nm (2.8)
In other words, solely the Ru*(bpy)32+ molecules that are in a close vicinity
(<59 nm) to the Ti0 2 surface are active for the reaction (2.7a); the rest only give rise to
a thermal dissipation of the absorbed light in the solution. To minimize this parasitic
effect, Ru(bpy)32+ was confined in a thin film of the cation exchange fluoropolymer,
Nation (cf. Table 2.6) [2.45].
In 1979, Goodenough et al. [2.58] replaced Ru(bpy)32+ by another complex
Ru(bpy)2L2+; L = 2,2/-bipyridine-4,4/-dicarboxylic acid, which adsorbed on to the
Ti0 2 single crystal [2.58]. In 1985, Graetzel et al. [2.57] pioneered the use of poly-
crystalline titania, and sensitized it with another derivative of Ru-bipyridine, viz.
RuL32+. These research efforts have provoked many follow-up studies [2.59-2.65],
54 Dye-Sensitized Solar Cells

Table 2.6 Examples of liquid-junction solar cells employing a sensitized Ti0 2 as the
photoanode. L = 2,2/-bipyridine-4,4/-dicarboxylate, L1 = 4,4/,4/-tricarboxy-2,2/:6/2/-terpyri-
dine, L2 = 4,4/-bis(5-hexylthiofen-2-yl)-2,2/-bipyridine, EC = ethylene carbonate, PC = pro-
pylene carbonate, AN = acetonitrile, MAN = methoxyacetonitrile, VN = valeronitrile, MPN =
methoxypropionitrile.

Ti0 2 Sensitizer Anax Electrolytea IPCE <%ol Ref.


photoanode [nm] [%] [%]

Single crystal Ru(SCN)2L2 530 I-/I3 + H 2 0 0.11 <l b [2.43]


anatase
Fractal/Ti Ru(bpy)32+ 452 HQ/Q + H 2 0 2.6 <l b [2.57]
(in solution)
Fractal/Ti Ru(bpy)32+ 452 HQ/Q + H 2 0 8 <l b [2.45]
(in Naflon)
Fractal/Ti RuL32+ 460 HQ/Q + H 2 0 44 «lb [2.57]
P25 + RuL2[//-(NC) 478 F/1- + EC/PC 80 -5b [2.47]
electrodep. Ru(CN)(bpy)2]2
Sol-gel + RuL2[//-(NC) 478 I-/I3 + EC/AN 85 7.12 [2.59]
electrodep. Ru(CN)(bpy)2]2
Sol-gel + TiCU RuCSCN^L1 610 I-/I3- + MAN + 80 10.4 [2.61,
DPI + PB 2.110]
Double layer Ru(SCN)2L2 530 I-/I3- 88 11.04 [2.65]
Double layer Ru(SCN)2LL2 580 I-/I3 + AN/VN 89 11.0 [2.178]
[C101]
Nanocryst. rutile Ru(SCN)2L2 530 I-/I3- + MPN 35 2.2 [2.124]
a
The electrolyte solution sometimes contains further additives such asteri-butylpyridine,ionic liquids,
etc. For the sake of simplicity, only the main solvent is quoted in the table.
b
Unpublished data.

but state-of-the-art DSCs still employ this original philosophy, i.e., the adsorption of
dyes (often Ru-bipy complexes) on titania [2.57]. For an optimum solar efficiency, the
Ti0 2 electrode must be sufficiently rough. The improvement in IPCE for a rough Ti0 2
surface can be demonstrated by a simple model calculation, assuming the extinction
coefficient (ε) of Ru(SCN)2L2 dye (at Amax = 530 nm) to be 1.27-107 cm2/mol and the
monolayer coverage of this dye to be Γ ~ 0.55 molecules/nm2. The IPCE of the sen­
sitized redox process is a product of quantum yield of charge injection, η{η^ and the
light-harvesting efficiency:

(2.9)

which gives (at ideal conditions, for 77inj = 100 %) a maximum theoretical IPCE of
0.27 % for a monolayer of Ru(SCN)2L2 on a flat surface. (The experimental value for
a single crystal was 0.11 %, see Table 2.6 and Section 2.4.5) [2.43]. The nanocrystal-
line electrode can be modeled as a layer (thickness δ) of close-packed spheres (diam­
eter d) deposited on a conducting support, such as FTO. Consequently:

(2.10)
Titania in diverse forms as substrates 55

If we fix (5= 10 ßm and 77inj = 100 %, we obtain IPCE = 69 % (for d = 100 nm)
and IPCE = 99 % (for d=10 nm). The modeled values (£= 10/mi d = 10 nm) are
thus close to the experimentally optimized ones [2.61]. A further improvement of
the surface roughness was achieved by electrodeposition (cf. Section 2.4.2) of small
anatase crystals [2.47, 2.49] on top of the "ordinary" anatase nanocrystals fabricated
by a sol-gel procedure (10-20 nm) [2.59] or from P25 [2.47]. In practical cells, the
blocking underlayers, prepared either by electrodeposition [2.47] or spray pyroly-
sis [2.50, 2.66], further enhance the cell performance by impeding the back electron
transfer (Eq. 2.7c). This issue is particularly important for solid-state solar cells [2.60,
2.61,2.67,2.68].
Although IPCE is a useful parameter for the evaluation of solar cells, a more
practical criterion is the total solar conversion efficiency, Φ8θ1, defined as the max­
imum electric power output, or in other words as the short circuit current density
(7Sc) x open-circuit-voltage (Uoc) divided by the white light solar power density, Psol
(for AMI.5, the conventionally used value of Psol is 1000 W/m2):

(2.11)

The parameter FF is the so-called fill factor describing the non-ideality of the
particular current/voltage profile:

(2.12)

with / being the photocurrent density, and U the voltage for which the Φ80ι is maximal.
There are various other practical issues to be addressed in real solar cells, such as the
cell geometry [2.69], the light intensity [2.70] and heat losses, including the effect
of wind [2.71]. Table 2.6 compiles several examples of solar cells; all comprising
the classical liquid-junction solar cells. An alternative design is based on replacing
the liquid electrolyte by a solid hole-conducting medium, such as CuSCN or spiro-
bis-fluorene derivatives [2.60, 2.61]. The efficiency of the latter is not yet competitive
[2.60, 2.61], but they are advantageous when it comes to stability and easy handling,
e.g., with plastic substrates.

2.3 ELECTROCHEMISTRY OF TITANIA: ACCUMULATION


REGIME

The depletion regime of Ti0 2 is generally studied through photoanodic reactions


(Section 2.2). On the other hand, the accumulation of electrons in the space-charge
layer can also be investigated electrochemically in the dark. Examples of faradaic
processes on Ti0 2 electrodes in the dark include: (i) the oxidation of Ti(OH)2+, leading
to an electrodeposition of Ti0 2 (Section 2.4.2) [2.47], (ii) water reduction, applicable
to estimate ΕΆ (Section 2.4.5) [2.72], (iii) electrochemical n-doping (Section 2.4.5)
[2.73], (iv) cathodic decomposition of propylene carbonate (Section 2.3.3) [2.54] and
56 Dye-Sensitized Solar Cells

(v) Li-insertion (Section 2.3.2). The effect of lithium insertion is particularly impor­
tant for understanding the performance of DSC [2.74-2.76].

2.3.1 Capacitive processes


The obvious process in the accumulation regime is double-layer charging at potentials
below Eft. This can be simply monitored by cyclic voltammetry [2.54]. The voltam-
metric current (/) of double layer charging scales with the first power of scan rate (v)
according to:

(2.13)

where Q is the voltammetric charge and C is the capacitance. The latter is typically
4-15 /iF/cm 2 in aprotic solutions containing Na + , K+, Cs + and NBu 4 + (Bu = butyl), but
increases by one order of magnitude in aqueous media, due to the H+/OH~ exchange
reactions in surface hydroxyls [2.54].
In the presence of Li + , the charge accumulation mechanism differs from the
capacitive process (Eq. 2.13), since both anatase [2.43, 2.54, 2.72, 2.77-2.87] and
rutile [2.43, 2.88] accommodate Li + also in the bulk crystal (cf. Section 2.3.2). The
monoclinic Ti0 2 (B) exhibits an unusual behavior (Figure 2.7) [2.44], consisting in
the Li-storage in Ti0 2 (B) formally obeying Eq. 2.13, but the charge exceeding that of

Fig. 2.7 Left panel: Cyclic voltammograms of fibrous titania made from Na-titanates. It was
composed of a mixure of Ti02(B) + anatase. Electrolyte solution: 1M LiN(CF3S02)2 + ethylene
carbonate/dimethoxyethane; scan rate 0.1-1.0 mV/s (in 0.1 mV/s steps for plots from bot­
tom to top). The inset shows a scanning electron microscopy image of the studied sample.
Right panel: Normalized peak current, z/%, where % is the peak current at the slowest scan
rate (0.1 mV/s). Circles: anodic peak at ca. 2 V (anatase); crosses: cathodic peak at ca. 1.6 V
(Ti02(B)), squares:cathodic peak at ca. 1.5 V (Ti02(B)). Reprinted with permission from Chem.
Mater. 17, 1248 (2005). Copyright 2005 American Chemical Society.
Titania in diverse forms as substrates 57

double-layer capacity. It is, actually, comparable to, or even higher than, the charge
corresponding to the bulk Li+ insertion in the Ti0 2 (anatase, rutile) lattice. An attrac­
tive issue is that the charging of Ti02(B) is not kinetically controlled by solid-state
diffusion (Section 2.3.2) [2.44]. This is demonstrated in Figure 2.7, which shows a
cyclic voltammogram of Li-insertion into a mixture of Ti02(B) and anatase. The ana­
tase component obeys the usual diffusion-controlled electrochemistry of Li-insertion
(see Figure 2.7, Section 2.3.2 and Eq. 2.16 below). The peculiar pseudocapacitive Li-
storage in Ti02(B) was discussed in terms of the channel-structure of fiber (Figure 2.2),
and two effects were considered: (i) a fast Li+ transport in the open channels running
parallel to the b-axis (fiber axis, cf. Figure 2.2) and (ii) a reversible trapping of Li+
ions at two sites in the lattice [2.44]. Both effects need to occur simultaneously, and
this is indirectly evidenced by the fact that two other channel titania structures, i.e.,
Ti0 2 (ramsdellite) and Ti0 2 (hollandite) do not seem to show this pseudocapacitive
Li-storage [2.44].

2.3.2 Li-insertion electrochemistry

The significance of Li+ in the electrolyte solution of a DSC is highlighted by the fact
that Li+ is potential-determining in aprotic solvents [2.89] and is commonly present
in electrolyte media for practical devices. The Li +/Ti02 interaction ranges from spe­
cific adsorption to insertion into the bulk crystal; the latter sometimes wrongly being
termed 'intercalation'. (In fact, the term 'intercalation' is reserved for laminar host
structures [2.90], which is not the case for titania). The insertion/adsorption of Li+ in
Ti0 2 causes two effects: (i) a positive shift of the conduction band edge and (ii) the
creation of deeper electron-trapping states, which play a role in electron transport
through Ti0 2 to the current collector [2.74-2.76].
Various forms of titania accommodate lithium electrochemically, which can be
schematically described by the equation:
(2.14)
Besides influencing the band energetic and electron transport in Ti0 2 photoan-
odes for DSC (see above), this reaction finds an application in Li-batteries [2.77, 2.78,
2.91]. Although most research reports concern anatase, Li-insertion into nanocrystal-
line rutile has recently attracted considerable attention [2.92, 2.93].
The accommodation of lithium in anatase is controlled by Li+ diffusion between
two-intermixed host phases (I\xlamd and Immd) [2.94, 2.95]. The advantage of a
nanocrystalline host for Li-insertion batteries can be documented by simply estimating
the diffusion layer thickness (dO) for Li+ underneath the Ti0 2 surface according to:
(2.15)
+ 13 2
For the solid state diffusion coefficient of Li in anatase, Du ~ 10~ cm /s (see
Table 2.4), the charging time, t, equals approx. 300 s if dO = 100 nm, whereas t ~
3 s if dO = 10 nm. In other words, an anatase crystal of dimensions (2 x dO) = 20 nm
becomes fully loaded with Li+ within a few seconds, whereas a ten times larger crystal
requires 100 times longer for a complete charging with Li+. Although such a model
calculation neglects the dependence of Du on the particle size [2.96], the benefits of a
58 Dye-Sensitized Solar Cells

nanocrystalline Li-insertion host are demonstrated. In this context, we may recall the
discussion in Section 2.2.3 (Eq. 2.10), where we came to a formally identical conclu­
sion regarding the improvement of the solar cell efficiency with a decreasing Ti0 2
particle size. Although the physical backgrounds for both conclusions are dissimilar
(Eqs. 2.10, 2.15), the gain in performance of both devices, i.e., solar cells as well as
Li-ion batteries employing nanocrystalline Ti0 2 , is obvious.
The insertion of Li into Ti(IV) oxides is suitable for addressing fundamental
academic questions. In cyclic voltammetry, the insertion/extraction peak current (/p)
scales with the square root of the scan rate, v [2.72, 2.83, 2.971 according to:

(2.16)

where n is number of electrons, A is the electrode area, c is the maximum concentra­


tion of Li+ in the accumulation layer (0.024 mol/cm3 for x = 0.5; cf. Eq. 2.14), a is
the charge-transfer coefficient, and the remaining symbols have their usual meaning.
In potential-step chronoamperometry, the insertion/extraction current (/) obeys the
Cottrell equation [2.43, 2.72, 2.82, 2.83, 2.96]:

(2.17)

Eqs. (2.16, 2.17) provide two independent techniques for the determination of
Du. (Yet another technique is galvanostatic intermittent titration [2.85]). The most
accurate data were obtained for anatase single crystals [2.43, 2.72], and it was even
possible to trace an anisotropy in Li-transport normal to the (101) and (001) anatase
faces (Table 2.4). This can be interpreted in terms of the Li+ hopping between pseudo-
octahedral positions in the anatase lattice.
In order to assess the quality of the Li-insertion host, we can also determine the
standard rate constant of charge-transfer, k0. The voltammetric peak current, /p, scales
with the peak potential, Ep [2.72, 2.82, 2.97] according to:

(2.18)

where Ef is the formal potential of insertion/extraction. Also in this case, we observe


the same trend in Li-insertion rate at the (101) and (001) faces of anatase (Table 2.4).
The matching of voltammetric and chronoamperometric values of Du is poorer
for nanocrystalline anatase [2.83], but Du decreases by several orders of magni­
tude. Analogously, the rate constants, k0 (Eq. 2.18) decrease for anatase nanocrystals
by orders of magnitude [2.82, 2.97] as compared to the values of the single crystal
(Table 2.4). A slower Li-insertion was also found for nanocrystalline Li4Ti5Oi2 (spi­
nel) with a monotonous drop of Du for a decreasing particle size [2.96]. However, the
sluggish Li-storage in nanocrystals need not be detrimental for the overall charging
rate of the electrode [2.82, 2.83, 2.96]. The reason for this is that a slower Li-diffusion
and smaller rate constants for nanocrystals (both Ti0 2 [2.82, 2.83] and Li4Ti5Oi2
[2.96]), are more than compensated by the increase of the active electrode area [2.82,
2.83, 2.96]. In other words, the bulk of solid is less important in small nanocrystals.
They virtually extrapolate to a 2-D system, exhibiting exclusively surface-confined
Titania in diverse forms as substrates 59

(pseudocapacitive) Li storage [2.96]. (For a further discussion of pseudocapacitive


Li-storage see Section 2.3.1 and [2.44]).
The effects associated with Li-insertion are particularly useful for investigat­
ing organized Ti0 2 materials [2.82-2.84, 2.86, 2.87]. The Li-insertion electrochem­
istry of anatase inverse opal evidenced that the charge propagation was hindered
due to a limited inter-particle necking [2.87]. This needs to be considered for the
application of inverse opal in high-drain devices such as solar cells [2.98], where cur­
rents of ca. 20 mA/cm2 are required for a 10 % solar cell at full sun illumination (cf.
Section 2.2.3).
The surfactant-templated mesoporous anatase, stabilized by Zr [2.82], Al [2.97],
as well as its non-stabilized counterpart [2.83, 2.84], exhibits specific voltammetric
features at potentials between ca. 1.3 and 1.6 V [2.79]. These are missing in ordinary
anatase, and have been denoted "S-peaks" (where the term "A-peaks" corresponds to
ordinary anatase) [2.83]. The name comes from the fact that they behave like surface-
confined redox couples. This is documented by a very small (Ev-Ef) splitting (even
smaller than 30 mV expected for reversible diffusion-controlled process, Eqs. 2.16,
2.17) and by the capacitive-like I(v) dependence (Eq. 2.13). The S-peaks also occur
in nanocomposites of Ti0 2 -P 2 0 5 [2.99], nanorods [2.100] nanotubes [2.101, 2.102]
and inverse opal [2.87]. They were first assigned to amorphous titania in the organ­
ized mesoporous skeleton [2.83], but further studies of titania nanosheets [2.86] and
nanofibers [2.44] have provided arguments for a different interpretation. The pseudo­
capacitive Li-storage can be reasonably assumed in nanosheets, exhibiting even 2-D
quantum confinement effects (cf. Eq. (2.4)) [2.86] and nanofibrous Ti02(B), which
also behaves like a nanosheet [2.44]. Eventually, it was found that the S-peaks evi­
denced the presence of Ti02(B) in all studied materials [2.44, 2.83, 2.84, 2.86, 2.87],
although this logical link was not considered prior to our detailed work on Ti02(B)
[2.44, 2.87, 2.99-2.102]. The obvious reason is that Ti02(B) may easily go unno­
ticed in a mixture with large concentrations of anatase. This is due to an overlap of
the diagnostic XRD peaks, especially in nanocrystalline materials with broad diffrac­
tion maxima [2.44]. Nevertheless, this conclusion demonstrates an excellent analyti­
cal sensitivity of Li-insertion electrochemistry for a mixture of nanocrystalline Ti0 2
phases (cf. Fig. 2.7). Other examples of the unique electrochemical "phase analysis"
is the detection of trace rutile in anatase [2.82] and trace anatase in Li4Ti5Oi2 [2.96,
2.103,2.104].

2.3.3 Spectroelectrochemistry of titania in the accumulation regime


Thin layers of Ti0 2 deposited on FTO are optically transparent, rendering them use­
ful for in-situ optical (UV-Vis) transmission spectroelectrochemistry. Particularly
valuable is the measurement at potentials near the ΕΆ, during which a broad band
of conduction-band electrons appears at 600-650 nm [2.48-2.50, 2.89]. This further
causes bleaching of interband transitions near the absorption threshold (Burnstein
shift) [2.48-2.50]. Both effects are applicable for determining of ΕΆ in nanocrystalline
materials, where standard impedance techniques often fail due to frequency disper­
sion and non-linear Mott-Schottky plots [2.49, 2.50]. Nevertheless, certain materials,
such as Ti0 2 prepared through aerosol pyrolysis (Section 2.2.3), allow a comparison
60 Dye-Sensitized Solar Cells

of both techniques [2.50]. The spectroelectrochemical method evidenced a negative


shift of £fb in quantum-sized anatase crystals obtained by electrodeposition (the shift
of £fb was ca. 0.2 V as compared to the "normal"-size anatase) [2.49]. An almost
identical ΕΆ shift was confirmed in Ti0 2 nanosheets, where the Q-size effect influ­
enced only one coordinate (cf. Eq. 2.4) [2.48].
In-situ Raman spectroelectrochemistry on single crystal anatase turned out to
be a useful technique for evaluating the course of electrochemical doping [2.73]. In-
situ infrared spectroelectrochemistry of nanocrystalline Ti0 2 was applied in order
to monitor the cathodic breakdown process in a solution of L1CF3SO3 + propylene
carbonate [2.54]. The solvent was found to decompose below 1.2 V vs. Li/Li+ in the
presence of trace water:

(2.19)

2.4 TITANIA PHOTOANODE FOR DYE SENSITIZED


SOLAR CELLS

Since the pioneering study in 1972 [2.46], titanium dioxide has been recognized as
an important material for the fabrication of photoanodes in photoelectrochemical and
solar cells [2.43, 2.45-2.47, 2.50, 2.57-2.59, 2.61, 2.64]. Thanks to its low price, stabil­
ity and environmental compatibility, Ti0 2 is attractive for practical applications. For
solar energy conversion, on the other hand Ti0 2 needs to be sensitized by suitable dyes
[2.43, 2.45,2.47,2.57, 2.58,2.61, 2.64]. The first photoelectrochemical studies of Ti0 2
in the early 70s employed almost exclusively the n-doped rutile single crystal as a pho­
toanode. This electrode material was unsuitable for sensitization due to the monolayer
of adsorbed dye on a flat surface not absorbing the light very efficiently (cf. Section
2.2.3, Eq. 2.9) [2.43, 2.45, 2.58]. To increase the physical surface area per square unit
of the projected electrode area (roughness factor), polycrystalline electrodes need to be
employed [2.45, 2.47, 2.57]. A logical conclusion is that small crystals (nanocrystals)
are optimum materials for Ti0 2 electrodes in solar cells. Also for other electrochemical
applications, such as electrochromic displays [2.105, 2.106] and Li-ion batteries [2.77-
2.79, 2.82, 2.91], the nanocrystalline morphology is beneficial for fast electrochemical
charging/discharging. Obviously, there is a motivation for exploring of electrochemical
properties of nanocrystals in general, and titania in particular.
Porous titania is a unique electrode material, due to its very low inherent con­
ductivity. Moreover, the nanocrystals do not support a built-in electrical field. The
charge transport is modeled by electron detrapping from states below the conduction
band edge [2.107]. In real DSC, the titania film is interpenetrated with an electrolyte
(solution), hence it actually behaves like a 3D-eletrochemical interface. The inter-
facial structure can be tuned by materials engineering, which gives rise to various
Titania in diverse forms as substrates 61

fundamental and practical questions. This chapter reviews the synthesis of titania
nanocrystals as well as the relations between the morphology of titania and the effi­
ciency of DSC.

2.4.1 Non-organized titania made by decomposition of Ti(IV)


alkoxides
As mentioned above, nanocrystals with sizes down to 10-20 nm, tend to grow in the
anatase structure. Consequently, nanocrystalline rutile is less common [2.88, 2.92]
than nanocrystalline anatase. A generic reaction for production of nanocrystalline
anatase is the hydrolysis of titanium(IV) alkoxides:
(2.20)

where R is an alkyl; typically an ethyl, butyl or isopropyl group. This reaction is the
basis of various sol-gel processes [2.36, 2.108]. The nanocrystals are further proc­
essed by the peptization (de-agglomeration), hydrothermal growth (Ostwald ripen­
ing) and sintering into layers (ca. 10 ßm in thickness) of nanoparticles (10-20 nm in
size) on certain substrate, such as FTO [2.54, 2.59, 2.61, 2.79, 2.82] or Ti [2.54, 2.79,
2.82]. Figure 2.8 shows a typical morphology for this material, which turned out to be
particularly suitable for the fabrication of solar cells [2.59, 2.61].
The application of colloidal particles obtained by the reaction (2.20), followed
by their deposition on transparent conductive glass substrates (FTO), was pioneered
by O'Regan et al. [2.109]. Originally, Ι.Ί-μνα films were prepared by sintering 8-nm
colloidal particles on top of the FTO substrate. Barbe et al. [2.108] further optimized
the synthetic parameters, such as the type of precursor (R in Eq. 2.20), peptization
(either in acidic or alkaline media, cf. also Section 2.4.4), hydrothermal growth,
binder addition and sintering conditions. The best electrode material of this type was
made from a mixture of two kinds of colloidal particles (peptized in acidic or alkaline
media, respectively) and further impregnated with TiCl4 (cf.Section 2.4.7) [2.110].

Fig. 2.8 A scanning electron microscopy image of a nanocrystalline anatase electrode. The
titania material was prepared by hydrolysis of titanium(IV) isopropoxide, peptized in an acidic
medium, autoclaved at 240 °C, deposited on FTO by doctor blading and sintered at 450 °C for
30 minutes.
62 Dye-Sensitized Solar Cells

The optimization of particle size upon hydrothermal growth was investigated by


Cao et al. [2.111]. The films were deposited by doctor blading or by screen-printing
from viscous pastes prepared by the aid of additives. Alternatively, the electro-
phoretic deposition from surfactant-stabilized colloids can also be used for DSC
applications [2.112].
A simplified synthetic protocol was adopted in earlier studies on DSCs: the
anatase layers were grown by repeated thermal decomposition of Ti(OR)4 on top of
a Ti-metal [2.45, 2.57, 2.113, 2.114] or FTO substrates [2.115]. These films were
termed 'fractal anatase layers', which reflects the morphology revealed by electro­
chemical impedance [2.113]. By repeated layer-by-layer deposition of 10 layers, a
film thickness of ca. 20 ßm was achieved with a roughness factor of ca. 200. The
roughness factor is defined as the ratio of the physical surface area of the poly-
crystalline film and the projected electrode area. The physical surface area is most
simply determined by N2 or Kr adsorption isotherms (BET method) [2.79], but also
the adsorption of the actual dye can be used. For instance, the pioneering work on
fractal titania electrodes employed the adsorption of RuL43" (L = 2,2'-bipyridine-4,4'-
dicarboxylate) assuming that one molecule occupies 1 nm2 at the monolayer cover­
age [2.113]. It is obvious that the second method provides the lower limit of the
roughness factor, since small pores (micropores) may not be accessible to the dye
molecule [2.113].
The hydrolysis of alkoxides (Eq. 2.20) can provide more complex materials,
such as Zr-doped anatase (1-5 % Zr) [2.82]. There have been numerous modifications
of the model sol-gel process depicted in Eq. (2.20). The hydrolysis of a mixture of
Ti- and Li-alkoxides may also lead to a useful ternary oxide, Li4Ti5Oi2 (spinel) [2.96,
2.103,2.104]:
(2.21)

where R = isopropyl or butyl and Et = ethyl. The usual product of reaction (2.21) is a
mixture of anatase with lithium titanates, but a proper tuning of the synthetic condi­
tions leads to a phase-pure spinel [2.104].

2.4.2 Electrochemical deposition of titania


Another preparative technique towards nanocrystalline anatase is based on anodic
oxidative hydrolysis of aqueous solution of TiCl3 [2.47]. At a pH between 2 to 2.5,
the reaction starts from partly hydrolyzed species, TiOH2+ (Eq. 2.22a). This is oxi­
dized on a Ti0 2 surface, and further hydrolyzed providing Ti0 2 (anatase) as the final
product (Eq. 2.22b).

(2.22a)

(2.22b)

Since Ti0 2 grows on the previously deposited Ti0 2 surface, the faradaic process
can only occur via an electron injection from TiOH2+into the conduction band of Ti0 2
[2.47]. Anodic reactions on (non-illuminated) Ti0 2 are, apparently, impossible above
Next Page

Titania in diverse forms as substrates 63

the flatband potential. Rather, they are restricted to redox couples with sufficiently
negative redox potentials [2.47].
The described electrodeposition of Ti0 2 offers various attractive features: (i) the
growth layer is well controlled by the passed charge (although the faradaic efficiency
is not 100 %, and is pH-dependent) [2.47]; (ii) the electrodeposited particles can be
extremely small, exhibiting even Q-size effects [2.49]; (iii) the particles are formed
from pure anatase, and can be efficiently sensitized for solar applications [2.47]; (iv)
the electrodeposited films thinner than ca. 200 nm are defect-free, rendering the film
applicable for blocking underlayers on F-doped Sn0 2 substrate [2.47]. The reactions
(2.22a, 2.22b) are also applicable for templated growth of anatase nanotubes [2.116],
mosaic nano-arrays [2.117], inverse opal [2.87] and gyroid network [2.118, 2.119].
The latter structure seems to be ideal for DSCs, particularly for solid-state solar cells
(See Section 2.4.4 for further discussion).
Another strategy of electrochemical growth of Ti0 2 for DSC was recently
developed by Oekermann et al. [2.120]. Titania was electrodeposited from a basic
solution of poly-Ti(IV)-oxo anions made from titanium isopropoxide and tetramethy-
lammonium hydroxide. The electrochemical process consisted in anodic oxidation of
hydroquinone to quinone on an FTO electrode, which produced protons. A lowering
of the pH caused precipitation of porous films on FTO which, upon calcination, con­
verted to anatase. The film with a roughness factor of 50 adsorbed N3 dye (5.16 nmol/
cm2) and exhibited a solar conversion efficiency of merely 0.2 %, presumably due to
residual impurities [2.120].

2.4.3 Aerosol pyrolysis


Nanocrystalline anatase can also be prepared by the pyrolysis of an aerosol from an
ethanolic solution of di-/1s,6>-propoxy-titanium-bis(acetylacetone) [2.50]. Similarly to
electrodeposition [2.47] this technique is suitable for fabricating blocking layers in
solid-state Ti0 2 solar cells [2.60]. The crucial effect consists in preventing the back
electron transfer from the supporting material (FTO) into the hole-transport medium,
which would short-circuit the cell [2.60, 2.61]. Peter et al. [2.66] demonstrated that
the blocking of the back electron transfer via aerosol-pyrolyzed underlayers is signifi­
cant also in liquid-junction solar cells, particularly at low light intensities. Moreover,
Burke et al. [2.68] reported that the blocking underlayer is important for DSCs using
organic ('Ru-free') dyes, which led to a conclusion that the Ru-based sensitizers
'insulate' the anode against recombination losses. Another attractive feature of spray
pyrolysis is its versatility; the doping of Ti0 2 by Fe or Nb can, for instance, be easily
carried out in this way [2.50].
The prepared layers generally present the anatase structure, with the size of
coherent crystal domains being around 50 nm [2.50]. Furthermore, they exhibit an
excellent photoelectrochemical performance of water oxidation [2.50]. The high pho-
toelectrochemical efficiency and easy fabrication of these Ti0 2 layers makes them
ideal for investigating various photoredox reactions [2.51-2.53] beyond the water
oxidation [2.50]. Cameron and Peter [2.121] presented a detailed treatment of elec­
trochemical and photoelectrochemical properties of aerosol-pyrolyzed Ti0 2 layers by
impedance measurements, ellipsometry and voltammetry.
CHAPTER 3

MOLECULAR ENGINEERING OF SENSITIZERS


FOR CONVERSION OF SOLAR ENERGY INTO
ELECTRICITY
Jun-ho Yum and Md. K. Nazeeruddin

3.1 INTRODUCTION

The greenhouse effect is related to the presence of carbon dioxide in the Earth's
atmosphere, along with many other pollutants, released by the extensive use of the
finite conventional energy supplies [3.1, 3.2]. The increasing global need for energy
and the depletion of fossil fuel reserves necessitate the development of clean alterna­
tive energy sources [3.3, 3.4]. Nuclear power was once regarded to offer a solution
for the increasing energy demands, but the storage of the nuclear waste has proven
too frightening for future generations, who cannot tolerate the risk of radioactive fis­
sion products. Therefore, to power our planet we shall have to develop methods for
harnessing solar energy, which is clean, non-hazardous and infinite.
Among the approaches for harnessing solar energy and converting it into elec­
tricity, the dye-sensitized solar cell (DSC) represents one of the most promising meth­
ods for future large-scale power production from renewable energy sources [3.2-3.9].
In these cells, the sensitizer is one of the key components, harvesting the solar radia­
tion and converting it into electric current. The electrochemical, photophysical, and
molecular electronic properties of the sensitizer play an important role for charge-
transfer dynamics at the semiconductor interface. Over the last 20 years, ruthenium
complexes endowed with thiocyanate ligands have maintained a clear lead in per­
formance amongst thousands of dyes that have been scrutinized. Their validated effi­
ciency record under standard (air mass 1.5) reporting conditions stands presently at
10.4 ±0.3% [3.10]. The choice of ruthenium metal is of special interest for a number
of reasons: because of its octahedral geometry one can introduce specific ligands in a
controlled manner; the photophysical, photochemical and the electrochemical proper­
ties of these complexes can be tuned in a predictable way; and the ruthenium metal
possess stable and accessible oxidation states from I to IV. The goal of this chapter is
84 Dye-Sensitized Solar Cells

to elucidate the basic requirements of sensitizers and their application in dye sensi­
tized solar cells.

3.2 RUTHENIUM SENSITIZERS

The optimal sensitizer for the dye-sensitized solar cell should be panchromatic, i.e.
absorb visible light of all colors. Ideally, all photons below a threshold wavelength
of about 920 nm should be harvested and converted into electric current. This limit is
derived from thermodynamic considerations, showing that the conversion efficiency
of any single-junction photovoltaic solar converter peaks at approximately 33 per­
cent near a threshold energy of 1.4 eV [3.11, 3.12]. The other essential property
required for the light-harvesting system of a molecular/semiconductor junction is that
the sensitizer in the excited state possess directionality. This directionality should
be engineered to provide an efficient electron transfer from the excited dye to the
Ti0 2 conduction band via good electronic coupling between the lowest unoccupied
molecular orbital (LUMO) of the sensitizer and the Ti 3d orbital. Also, the sensitizer
should have suitable interlocking groups for grafting the dye on the semiconductor
surface to ensure intimate electronic coupling between its excited state wave function
and the conduction-band manifold of the semiconductor [3.5, 3.13].
The spectral properties of ruthenium sensitizers can be tuned by introducing a
ligand with a low-lying π * molecular orbital and by destabilization of metal t2g orbital
through the introduction of a strong donor ligand. However, the extension of the
spectral response into the near-infrared region, at the expense of shifting the LUMO
orbital to lower levels from where charge injection into the Ti0 2 conduction band
can no longer occur, is not useful [3.14]. On the other hand, near-infrared response
by destabilization of Ru t2g (HOMO) levels close to the redox potential of the redox
couple is also not useful because of problems associated with regeneration of the dye
by electron donation from iodide following charge injection into the Ti0 2 . Therefore,
the optimum ruthenium sensitizers should exhibit an excited-state oxidation potential
of at least -0.9 V vs. SCE in order to inject electrons efficiently into the Ti0 2 conduc­
tion band [3.8]. The ground state oxidation potential should be about 0.5 V vs. SCE,
in order to be regenerated rapidly via electron donation from the electrolyte (iodide/
triiodide redox system or a hole conductor). A significant decrease in electron injec­
tion efficiencies will occur if the excited- and ground-state redox potentials are lower
than these values. Various ruthenium(II) complexes are primarily employed as sen­
sitizers, with anchoring groups such as carboxylic acid, dihydroxy, and phosphonic
acid on the pyridine ligands. The remarkable performance of the tetraprotonated
[c/1s,-(dithiocyanato)-Ru-/7/1s,(2,2-bipyridine-4,4-dicarboxylate)] complex (1) and its
doubly deprotonated analogue, complex (2), see below, have had a central role in sig­
nificantly advancing DSC technology. The power conversion efficiency of a solar cell
is a product of the short-circuit photocurrent (/sc), the open-circuit potential (Voc) and
the fill factor, divided by the light intensity. In order to obtain high overall light-to-
electric-power conversion efficiencies, optimization of the short-circuit photocurrent
(/sc) and the open-circuit potential ( Voc) of the solar cell is essential.
Molecular engineering of sensitizers 85

3.2.1 Effect of protons carried by the sensitizers on the performance


The conduction band of the Ti0 2 is known to have a Nernstian dependence on pH [3.15,
3.16]. The fully protonated sensitizer (1), upon adsorption, transfers most of its protons
to the Ti0 2 surface, charging it positively. The electric field associated with the sur­
face dipole generated in this fashion enhances the adsorption of the anionic ruthenium
complex and assists electron injection from the excited state of the sensitizer into the
titania conduction band, favoring high photocurrents (18-19 mA/cm2). However, the
open-circuit potential (0.6 V) is lower due to the positive shift of the conduction band
edge induced by the surface protonation.
On the other hand, the sensitizer (3) that carries no protons shows high open-
circuit potential compared to complex (1), due to the relative negative shift of the
conduction-band edge induced by the adsorption of the anionic complex, though as a
consequence the short-circuit photocurrent is lower. Thus, there should be an optimal
degree of protonation of the sensitizer in order to maximize the product of short-
circuit photocurrent and open-circuit potential, which, along with the fill factor, deter­
mines the power-conversion efficiency of the cell.
The performance of the three sensitizers (1), (2) and (3) that contain different
degrees of protonation were studied on nanocrystalline Ti0 2 electrodes [3.13, 3.17].
Figure 3.1 shows the photocurrent-action spectra obtained with a monolayer of these
complexes coated on Ti0 2 films.

The incident monochromatic photon-to-current conversion efficiency (IPCE)


is plotted as a function of excitation wavelength. The IPCE value in the plateau
region is 80% for complex (1), while for complex (3) it is only about 66%. In the red
region, the difference is even more pronounced. Thus, at 700 nm, the IPCE value is
twice as high for the fully protonated complex (1) as compared to the deprotonated
complex (3). As a consequence the short-circuit photocurrent is 18-19 mA/cm2 for
complex (1), while it is only about 12-13 mA/cm2 for complex (3). However, it is
impressive that the photovoltage for complex (3) is 900 mV, which is 300 mV higher
than that for complex (1), as shown in Table 3.2. The increased open-circuit poten­
tial in the former case is the result of negative shift of the Fermi level which results
in an increase in the gap between the redox couple iodide/triiodide and the Fermi
level. Nevertheless, this is insufficient to compensate for the current loss. Hence, the
photovoltaic performance of complex (2), carrying two protons, is superior to that
86 Dye-Sensitized Solar Cells

Fig. 3.1 Photocurrent-action spectra of nanocrystalline Ti02 films sensitized by complexes


(1), (2) and (3). The incident photon to current conversion efficiency is plotted as a function of
wavelength.

Table 3.2 Photovoltaic performances based on different degrees of protons in sensitizers.

Sensitizer No. of /sc (mA/cm2) Voc (mV) Fill factor Efficiency


protons at AM 1.5
Complex (1) 4 19 + 0.5 600 ± 30 0.65 ± 0.05 7.4
Complex (2) 2 17 ± 0.5 730 + 30 0.68 + 0.05 8.4
Complex [TBA] (1)3 1 16.8 ± 0.5 770 ± 30 0.72 ± 0.05 9.3
Complex (3) 0 13 + 0.5 900 ± 30 0.7 ± 0.05 8.2

of compounds (1) and (3) that contain four or no protons, respectively. The doubly
protonated form of the complex is therefore preferred over the other two sensitizers
for sensitization of nanocrystalline Ti0 2 films. The data presented in Table 3.2 sug­
gest that the one-proton dye is the optimum for high-power conversion efficiency of
the cell.

3.2.2 Effect of cations in the ruthenium sensitizers on the performance


The anchoring groups of the sensitizer immobilize it on the nanocrystalline Ti0 2
surface, allowing high quantum yields of the excited-state electron-transfer process
[3.8, 3.17-3.22]. The anchoring group of the adsorbed dye transfers most of its pro­
tons/cations onto the semiconductor surface. Thus, to see the influence of cations
on the photovoltaic performance and stability, tetrabutylammonium cations were
replaced by sodium ions in complex (4).
Molecular engineering of sensitizers 87

Figure 3.3(a) shows the photocurrent density as a function of voltage of four


DSCs prepared with complex (2) and (4), both in combination with electrolytes E l
and E2 (see Table 3.4). The corresponding photovoltaic parameters of these fresh
cells (short-circuit current density, / s c ; open-circuit voltage, Voc; fill factor, FF and
efficiency, η) are presented in Table 3.5 as a function of light intensity. The highest
overall power-conversion efficiency was obtained for the system complex (2)/El elec­
trolyte (device C), due to the higher values of Voc and FF. However, the sodium-based
dye in combination with the same electrolyte (device A) presented the highest photo-
current density at all light intensities. Table 3.4 shows that the effect of cation sub­
stitution depends on the electrolyte composition. The methoxypropionitrile (MPN)
based electrolyte E l produces better efficiencies than E2, containing butyrolactone
(GBL) as a solvent, maximum efficiency values being 8.51 ± 0 . 0 1 % (0.1 sun) for
device A and 8.8 ± 0 . 1 % (0.1 sun) for device C.
The better performance of devices A and C, based on the E l electrolyte, is also
reflected in the incident photon-to-current conversion efficiency (IPCE) presented in
Figure 3.3(b). Devices A and C reach a maximum of approximately 72% at 530 nm
and 78% at 550 nm, respectively. On the other hand, the photocurrent-action spectra
for the E2-based systems show lower IPCE values at 530 nm: namely, a maximum of

Fig. 3.3 (a) Photocurrent intensity-voltage characteristics for devices A to D, measured at AM


1.5 (1000 W-m -2 ) global sunlight illumination, (b) Photocurrent action spectra for the same
devices.
88 Dye-Sensitized Solar Cells

Table 3.4 Systems under study, Electrolyte El contains 1.0 M l-Propyl-3-methylimidazolium


iodide, 0.15 M iodine, 0.1 M Guanidinium thiocyanate and 0.5 M N-butyl benzimidazole in
MPN. Electrolyte E2 contains 0.1 M iodine, 0.5 M N-methylbenzimidazole in a mixture of
BMILPMITFSLGBL (2:3:1) vol/vol.

Device Dye Electrolyte


A Complex (4) El
B Complex (4) E2
C Complex (2) El
D Complex (2) E2

Table 3.5 Detailed photovoltaic parameters under various sunlight intensity. The active area
of the cell with a metal mask was 0.158 cm2.
2
Device Light intensity 4/mA-cm Kc/mV FF ηΙ%

A 1 sun 15.10 ±0.01 777.0 ± 1 0.660 ± 0.01 7.60 ±0.2


0.5 sun 8.15 + 0.04 759.7 ±0.8 0.700 ± 0.01 8.21 ±0.06
0.1 sun 1.51 ±0.01 710.3 ±0.6 0.754 ± 0.001 8.51 ±0.01
B 1 sun 12.75 + 0.07 773.0 ±15 0.600 ± 0.02 5.90 ±0.1
0.5 sun 7.00 ± 0.2 766.0 ± 20 0.670 ± 0.02 6.80 ±0.2
0.1 sun 1.33 ±0.04 702.0 ± 21 0.756 ± 0.004 7.39 ±0.06
C 1 sun 14.90 ±0.1 797.0 ± 6 0.712 ±0.005 8.45 ± 0.06
0.5 sun 8.00 ± 0.06 779.0 ± 5 0.750 ± 0.004 8.80 ±0.08
0.1 sun 1.47 ±0.01 728.0 ± 5 0.783 ± 0.001 8.80 ±0.1
D 1 sun 13.00 ±0.01 767.0 ± 4 0.740 ± 0.01 7.35 ±0.08
0.5 sun 7.01 ±0.01 746.0 ± 5 0.760 ± 0.01 7.50 ±0.1
0.1 sun 1.29 ±0.02 685.0 ± 5 0.770 ± 0.01 7.30 ±0.01

54% for device B and 67% for device D. The highest IPCE obtained with device C is
therefore in line with its better photovoltaic performance.

3.2.3 Device stability


Besides efficiency, the long-term stability is also a key requirement for applications.
The above four cells were subjected to light soaking at 50°C for 1000 h. The evolu­
tion of the photovoltaic performance throughout the aging process is presented in
Figures 3.6 and 3.7. Device A, employing sensitizer complex (4) along with the El
electrolyte, shows an excellent stability, with a drop of less than 1 % in efficiency
(7.59% to 7.54%) during 1000 h aging. The small decrease in the / sc was compen­
sated by an increase in FF and Voc. By contrast device C suffered a decrease in the
power-conversion efficiency during the aging process from 8.40% to 7.11%. A drop
of all the photovoltaic parameters occurred during the aging time. Thus in spite of
the excellent initial performance exhibited by this system, it showed poor stability,
as observed in previous studies [3.23]. Moreover, combining the complex (2) with
Molecular engineering of sensitizers 89

Fig. 3.6 Evolution of photovoltaic parameters for device A (A) and device C (■), aged under
one sunlight soaking at 50°C.

the E2 electrolyte renders the system even less stable. After 1000 h of light soaking,
there was a drop in efficiency of about 25%, ascribed to major losses in all photo­
voltaic parameters. The same behaviour was observed for device B, suggesting that
the E2 electrolyte renders these devices unstable, probably due to desorption of dye
in the electrolyte. The promising results in terms of performance/long-term stabil­
ity of complex-(4)-based DSCs in combination with the El electrolyte are shown in
Figure 3.6.

3.2.4 Effect of alkyl chains in the sensitizer on the performance

One of the problems in dye-sensitized solar cells is water-induced desorption of the


sensitizer from the Ti0 2 surface. Extensive efforts have been made in our laboratory
to overcome this problem by introducing hydrophobic properties in the ligands (com­
plexes 5-9). The absorption spectra of these complexes show broad features in the vis­
ible region and display maxima around 530 nm. The performance of these hydrophobic
90 Dye-Sensitized Solar Cells

Fig. 3.7 Evolution of photovoltaic parameters for device B (Δ) and device D (D), aged under
one sunlight soaking at 50°C.

complexes as charge-transfer photosensitizers in nanocrystalline Ti02-based solar


cells shows excellent stability towards water-induced desorption [3.24].

The rate of electron transport in a dye-sensitized solar cell is a major element


of the overall efficiency of the cell. The injected electrons into the conduction band
from the optically excited dye can traverse the Ti0 2 network and be collected at the
transparent conducting glass, or it can react with either the oxidized dye molecule
Molecular engineering of sensitizers 91

(step 5) or the oxidized redox couple (step 6) in Figure 3.8. The reaction (recombina­
tion) of injected electrons into the conduction band with the oxidized mediator gives
undesirable dark currents, reducing significantly the charge-collection efficiency, and
thereby decreasing the total efficiency of the cell.
The surface treatment of Ti0 2 electrodes with organic molecules, such as pyrid-
ines [3.25,3.26] and co-adsorbent [3.27-3.32], has been shown to substantially improve
the photovoltage of DSCs due to retarded recombination. Mesoporous Ti0 2 electrodes
consisting of nanocrystalline particles with a thin overcoat of a different metal oxide
with a higher conduction-band edge, have been shown to minimize interfacial recom­
bination dynamics relative to unidirectional charge transport [3.33-3.35]. Gregg et al.
have examined surface passivation by deposition of insulating polymers [3.36]. The
hydrophobic sensitizers that contain long aliphatic chains (5-9) on Ti0 2 considerably
suppressed the recombination reaction. A significant retardation of the charge recom­
bination of injected electrons with oxidized sensitizers (step 6 in Fig. 3.8) is observed
with increasing alkyl-chain length up to complex (8), from -90 to 490/is [3.37]. In
terms of the electron-recombination process with oxidized redox media (step 5 in
Fig. 3.8), the complex-(9)-sensitized solar cell showed more than an order of magni­
tude slower recombination than the complex (5)-sensitized solar cell [3.37]. The most

Fig. 3.8 Illustration of the interfacial charge-transfer processes in nanocrystalline dye-sensitized


solar cells. S/S+/S* represents the sensitizer in the ground, oxidized and excited state, respec­
tively. Visible light absorption by the sensitizer (step 1) leads to an excited state, followed by
electron injection (step 2) into the conduction band of Ti02. The oxidized sensitizer (step 3) gets
reduced by Γ/IJ" redox couple (step 4). The injected electrons into the conduction band may
react either with the oxidized redox couple (step 5) or with oxidized dye molecule (step 6).
92 Dye-Sensitized Solar Cells

Fig. 3.9 Pictorial representation of the blocking of the oxidized redox couple lï from reaching
the surface of Ti02 for conduction band electrons using hydrophobic sensitizers, which form
an aliphatic network.

likely explanation for the reduced dark current is that the long chains of the sensitizer
interacts laterally to form an aliphatic network as shown in Figure 3.9, thereby prevent­
ing triiodide from reaching the Ti0 2 surface. Recently, the hydrophobic chains have
been incorporated to metal free organic sensitizer, leading to dramatically improved
stability [3.38-3.40]. However, it was apparent that increasing the chain length results
in a retardation of the dye regeneration reaction (step 4), hence complex-(9)-sensitized
solar cells are characterized by reduced device performance.

3.2.5 Effect of the ^-conjugation bridge between carboxylic acid


groups and the ruthenium chromophore
A new development for dyes as applied to solar cells comes from the preparation of
amphiphilic heteroleptic complex (7) in which the hydrophobic long chains interact
Molecular engineering of sensitizers 93

laterally to form an aliphatic network to impede the triiodide ions from reaching the
Ti0 2 surface [3.41]. To further improve the amphiphilic heteroleptic ruthenium sen­
sitizers, high molar-extinction-coefficient dyes were developed by incorporating the
^-conjugation bridge between anchoring group, carboxylate and ruthenium chromo-
phore. The complexes (10) and (11), by incorporating vinyl or phenylethenyl moie­
ties into the anchoring bipyridine, affords ruthenium sensitizers with increased molar
extinction coefficients and enhanced red response in the visible regions. Figure 3.10
shows the absorption spectra of complex (7), (10) and (11) sensitizers, which are
dominated by metal-to-ligand charge-transfer transitions (MLCT). The peak positions
of the lowest energy MLCT band of the complex (10) and (11) sensitizers are red
shifted by 15 nm and 18 nm, respectively, when compared with complex (7), and
the molar extinction coefficients of the new sensitizers are higher by 15% and 38%,
respectively. The increased ^-conjugation lengths of 4,4'-bis(carboxyvinyl)-2,2'-
bipyridine in complex (10) and 4,4/-bis((carboxyphenyl)ethenyl)-2,2/-bipyridine in
complex (11) are responsible for the increased molar extinction coefficients in the
visible region as compared with complex (7).

Figure 3.11 shows a comparison of the photocurrent-voltage (J-V) curves and


the incident-photon-to-current-conversion-efficiency (IPCE) spectra of the dye-
sensitized solar cells consisting of complex (7), (10) and (11) [3.42]. For the J-V
and IPCE measurements, we used Ti0 2 films consisting of a transparent 2.6-/mi-
thick mesoporous thin film to compare the effects of sensitizers in the dye-sensitized
solar cells (DSCs). The J-V data (Table 3.12) indicate that the photocurrents of
94 Dye-Sensitized Solar Cells

Fig. 3.10 UV-vis spectra of 1 x 10 5 M each of complex (7), (10) and (11) sensitizers in
EtOH.

Fig. 3.11 J-V curves of solar cells prepared with 2.6-/*m-thin Ti0 2 films impregnated with
complex (7) (dotted line), (10) (dashed line) and (11) (solid line) under 100 mW/cm2 of light
intensity. The inset shows the corresponding IPCE spectra.

complex-(lO)- and complex-(l l)-coated T i 0 2 cells are higher than those of complex
(7), whereas the open-circuit potentials are lower than for those of complex (7).
The short-circuit photocurrent density (/ sc ) of the T i 0 2 cell with complex (11) has
increased by 16% and the open-circuit voltage (Voc) of the same has decreased by
Molecular engineering of sensitizers 95

Table 3.12 Photovoltaic parameters of the dye-sensitized solar cells with different sensitizers
with 2.6-μηι thin Ti0 2 films under an illumination of AM-1.5 solar light (100 mW/cm2).

Sensitizer / sc [mA/cm2] Ko [mV] FF η [%]

Complex (7) 7.16 787.82 0.720 4.06


Complex (10) 7.80 738.23 0.719 4.14
Complex (11) 8.31 736.14 0.722 4.41

Fig. 3.13 Impedance spectra of complex (7), (10) and (11) DSCs based on 1.8 + 4μηι film
thickness measured under 1 sun at open-circuit potential, (a) Bode phase plots; (b) Nyquist
plots.

6.6% as compared with complex (7). As a result, the overall solar energy conver­
sion efficiency of the complex-(ll)-sensitized cell increased by 8.6%. And, the cells
consisting of complex (10) and (11) show consistently broad IPCE spectra compared
with that of complex (7). The enhanced IPCE values correspond to the increased
absorptions in the visible and near-infrared spectral region.
The Voc of a DSC is determined by the difference between the quasi-Fermi
level (EFn) in the Ti0 2 under illumination and the Fermi level of the electrolyte
(redox potential, EF>redox). The increase in Voc could be caused by two different
mechanisms, one is a retardation of the recombination between injected electrons
and oxidized species in the electrolyte, and another is band edge movement of
semiconductor with respect to the redox potential [3.28]. Electrochemical imped­
ance spectroscopy (EIS) and study of current and voltage decay provide valuable
information for the understanding and characterization of the basic photovoltaic
parameters of the DSCs. In Figure 3.13, the large semicircle in the Nyquist plot and
the characteristic frequency peak in the Bode phase plot arise from electron trans­
fer at the Ti02/electrolyte interface (step 5 in Fig. 3.8). The frequency peaks from
complex-(lO)- and complex-(ll) DSCs in the Bode phase plots (Figure 3.13(a))
are at higher frequencies than those from the complex-(7) DSC, indicating that the
electron lifetimes of the complex (10) and (11) are shorter than that of complex (7).
This impedance from complex-(lO) and complex-(ll) dye-sensitized solar cells
96 Dye-Sensitized Solar Cells

Fig. 3.14 A plot of apparent electron lifetime against photoinduced charge density of complex
(7), (10) and (11) DSCs with a 4.4 μτη Ti0 2 film thickness.

shows that the recombination process is faster than that for complex (7). And
apparent electron lifetime at fixed charge density obtained from photocurrent and
voltage decay in Figure 3.14 supports the impedance result [3.42]. This result can
presumably be correlated to the closeness of the carboxylic acid groups to the
ruthenium center in the chemical structure. Since complexes (10) and (11) have
distantly connected carboxylic acids to bipyridyl groups, the injected electrons can
be recombined with triiodide at the Ti02/electrolyte interface more readily than for
complex (7). In order to investigate a contribution of the band-edge shift to Voc, the
difference of the photoinduced charge density as a function of open-circuit voltage
was observed, and slightly positively shifted band edges were found in complex-
(10) and complex-(ll) sensitized solar cells. Hence the decreased Voc of these cells
can be explained by their faster recombination relative to that of complex (7) and by
the small influence of the band-edge shift.

3.2.6 High Molar Extinction Coefficient Sensitizers


To further improve the heteroleptic ruthenium sensitizers, high molar extinction coef­
ficient sensitizers (12-13) have developed by expanding the ^-conjugation of the
hydrophobic ligand and endowing it with electron donating alkoxy groups. In com­
plex (13), the purpose of 4,4/-di-(2-(3,6-dimethoxyphenyl)ethenyl)-2,2/-bipyridine
ligand that contains extended ^-conjugation with substituted methoxy groups is to
enhance molar extinction coefficient of the sensitizers, and furthermore to provide
directionality in the excited state by fine tuning the LUMO level of the ligand with
the electron-donating alkoxy groups. The absorption spectra of complexes (13) are
dominated by the metal-to-ligand charge transfer transitions in the visible region 400
and 545 nm. The molar extinction coefficients of these bands are close to 35,000 and
Molecular engineering of sensitizers 97

19,000 M xcm x, respectively which are higher (-40%) when compared to those of
the standard sensitizer, complex (2) (Figure 3.15).

Figure 3.16 shows an INDO/s and DFT study of the electronic and optical prop­
erties of complexes (2, 12, 13). The theoretical data suggest that the top three frontier
filled orbitals have essentially ruthenium 4d (t2g in octahedral group) character with a
sizable contribution coming from the NCS ligand orbitals [3.43, 3.44]. Most critically,
the calculations reveal that for the Ti02-bound sensitizers (12-13), excitation directs
charge into the carboxylbipyridine ligand bound to the Ti0 2 surface. Table 3.17 illus­
trates the comparison of IPCE and photovoltaic data, respectively obtained using
complex (2) and (13) with transparent Ti0 2 membrane of various thicknesses. The
data show that as the membrane thickness decreased, less photocurrent is produced
due to the fact that the incident light is not fully absorbed by the adsorbed dye. The
/ sc increases continuously with film thickness, reaching a plateau value of close to
19 mA/cm2 at 14 /mi while efficiency (77) reaches a maximum of 10.82%. The data
demonstrate that the photovoltaic performance of complex (13) is superior to that of
complex (2) for thinner membranes. The / sc of complex (13) is -30% higher than
the / sc of complex (2) on a 2-/mi- thick Ti0 2 membrane, consistent with the high
molar extinction coefficient of complex (13). However, for thicker Ti0 2 membranes
98 Dye-Sensitized Solar Cells

Fig. 3.15 Comparison of absorption spectra of complexes (2) (dotted line) and (13) (solid line)
in ethanol.

Fig. 3.16 Molecular orbital energy diagram of complexes (2), (12) and (13) compared to that of
a Ti0 2 nanoparticle model. HOMO-LUMO gaps (eV) and lowest TDDFT excitation energies
(eV, data in parentheses) are reported together with isodensity plots of the HOMO-3, HOMO
and LUMO of the complex (13).
Molecular engineering of sensitizers 99

Table 3.17 Comparison of photovoltaic parameters of complex (2) and (13) with different
transparent Ti02 membrane thickness using an electrolyte with the composition of 0.60 M
butylmethylimidazolium iodide (BMII), 0.03 M I2, 0.10 M guanidinium thiocyanate and
0.50 M tert-butylpyridine in a mixture of acetonitrile and valeronitrile (volume ratio of
85:15) under 99 mW/cm2 light intensity.
Ti02film IPCE ^sc Ko [mV] Fill Efficiency
thickness (/mi) Sensitizer (%) [mA/cm2] factor [%]
2 Complex (13) 61 9.3 809 0.76 5.72
Complex (2) 50 7.27 819 0.76 4.56
5 Complex (13) 74 12.12 783 0.76 7.31
Complex (2) 67 9.94 810 0.77 6.19
7 Complex (13) 77 13.42 782 0.75 8.04
Complex (2) 76 11.73 797 0.77 7.21
9 Complex (13) 76 13.94 772 0.77 8.31
Complex (2) 79 12.86 788 0.76 7.78

of 9 /mi, the disparity in efficiency between the two sensitizers decreases from 30% to
less than 7%. For thicker Ti0 2 membrane of 9 /mi and greater, an upper-bound limit
is reached, beyond which the influence of the high molar extinction coefficient of the
dye is buffered.

3.2.7 Tuning spectral response by thiocyanato ligands

The ideal sensitizer for a single-junction photovoltaic cell that converts standard
global AM-1.5 sunlight into electricity should absorb all light below a threshold
wavelength of about 920 nm. Indeed, the main drawback of complex (2) is the lack of
absorption in the red region of the visible spectrum. When engineering panchromatic
ruthenium complexes, the LUMO and HOMO of the complex have to be maintained
at levels where photoinduced electron transfer in the Ti0 2 conduction band, as well
as regeneration of the dye by iodide, can take place with practically 100% yield. The
spectral and redox properties of ruthenium polypyridyl complexes can be tuned in
two ways; first, by introducing a ligand with a low-lying ;r* molecular orbital; and
second by destabilization of the metal t2g orbital through the introduction of a strong
donor ligand. Heteroleptic complexes containing bidentate ligands with low-lying
7Γ* orbitals, together with others having strong σ-donating properties, demonstrate
impressive panchromatic absorption properties [3.45]. However, the extension of the
spectral response into the near-infrared has been gained at the expense of shifting the
LUMO orbital to levels that are too low to allow charge injection into the Ti0 2 con­
duction band [3.46, 3.47]. Near-infrared response can also be gained by the upward
shifting of the Ru t2g (HOMO) levels. However, it turns out that the mere introduction
of strong sigma-donor ligands into the complex often does not lead to the desired
spectral result as both the HOMO and LUMO are displaced in the same direction.
Furthermore, the HOMO position cannot be varied freely because the redox potential
of the dye must be maintained sufficiently positive to ensure rapid regeneration of
the dye by electron donation from iodide following charge injection into the Ti0 2 .
100 Dye-Sensitized Solar Cells

To fulfill the requirement of panchromatic ruthenium complexes, a complex (14) in


which the ruthenium center is coordinated to a monoprotonated tricarboxyterpyridine
ligand and three thiocyanate ligands has been synthesized [3.48].
Figure 3.18 shows the photocurrent-action spectrum of such a cell containing
complex (2) and (14), where the incident photon-to-current conversion efficiency is
plotted as a function of wavelength. It is evident that the response of the complex (14)
extends 100 nm further into the infrared than that of complex (2). The photocurrent
onset is close to 920 nm, i.e. near the optimal threshold for single-junction converters.
From there on the IPCE rises gradually until at 700 nm it reaches a plateau of over
80%. From the overlap integral of the curves in IPCE (Figure 3.18) with AM-1.5 solar
emission, one predicts the short circuit photocurrents of complex-(2)- and complex-
(14)-sensitized cells to be 16.5 and 20.5 mA/cm2 [3.49]. Routinely, the experimen­
tal photocurrents obtained with the complex (14) are in the range of 18-21 mA/cm2
[3.50]. The open circuit potential (Voc) is 720 mV, and the fill factor (FF) is 0.7, yield­
ing for the overall solar-to-electricity conversion efficiency a value of 10.4% (global
AM-1.5 solar irradiance 1000 W nr 2 ) [3.50]. With this complex (14), the conversion
efficiency of 11.1% was achieved using high haze Ti0 2 electrodes [3.51].

Fig. 3.18 Photocurrent-action spectra obtained with complex (2) (red line) and (14) (black
line) as sensitizer. The photocurrent response of bare Ti02filmsis also shown for comparison.
Molecular engineering of sensitizers 101

3.2.8 Non-thiocyanato ruthenium complexes


Complexes (2) and (14) endowed with thiocyanate ligands have become a paradigm
in the area of dye-sensitized nanocrystalline Ti0 2 films [3.52]. Their validated effi­
ciency record under standard air-mass-1.5 reporting conditions stands presently at
11% [3.51]. Many attempts to replace the thiocyanate donor ligands have been made,
since the mono-dentate SCN is believed to be the weakest part of the complex from
the chemical stability point of view. However, so far these efforts have yielded only
limited success, as the conversion efficiency achieved with complexes that do not
contain SCN remain well below few percent [3.53]. The absorption spectrum of
the complex (15) is dominated in the visible region by three bands at 406, 486 and
560 nm, which are attributed to metal-to-ligand (MLCT) charge-transfer transitions.
The lowest energy MLCT band is red shifted by 25 nm when compared to the
standard bis tetrabutylammonium c/1s,-dithiocyanatobis-2,2/-bipyridine-4-COOH,4/-
COO~ ruthenium(II), complex (2) [3.54]. The striking feature of complex (15) is
the presence of a new band at 486 nm, with a remarkably high molar extinction
coefficient, compared to complex (2) whose spectrum shows a valley in this wave­
length region. In theoretical calculations using both density functional theory (DFT)
for the ground state geometries and time dependent DFT (TDDFT) for the excited
state energies and properties, the computed spectrum in the visible region is in very
good agreement with the experimental one (Figure 3.19). The presence of carboxy-
lic groups allows stable anchoring of the sensitizer to the semiconductor surface,
so as to ensure high electronic coupling between the dye and the semiconductor,
which is required for efficient charge injection. The DFT calculations clearly show
that the optically active electronic excitations in the visible consist of transitions
from the three highest HOMOs, which lie within -0.35 eV and have a strong ruthe­
nium t2g character. The new band at 486 nm is characterized by a transition from the
HOMO, which has a sizable ;r-orbital contribution from the cyclometallated ligand,
as shown in Figure 3.20. The LUMOs involved in the former excitations are a set of
;r*-orbitals localized on the bipyridines ligands bearing the protonated carboxylic
groups. These orbitals will give rise to strongly mixed dye-Ti0 2 states upon absorp­
tion on the Ti0 2 nanoparticles. Therefore, the excited states generating the absorp­
tion spectrum might strongly couple to the Ti0 2 conduction band states, eventually
leading to efficient electron injection and increased photocurrents.
102 Dye-Sensitized Solar Cells

Fig. 3.19 UV-Vis experimental absorption spectra of complex (15) (red solid line) and com­
plex (2) (dashed blue line) measured in DMF solution. Red-bars represent the computed verti­
cal electronic excitations intensities (TDDFT) of complex (15). For three selected optically
active electronic transitions (black-bars with labels 1, 2 and 3 at 533 nm, 496 nm and 426 nm
respectively) the electron-hole density plots are shown (hole and excited electron densities are
represented in pink and cyan respectively).

The incident monochromatic photon-to-current conversion efficiency (IPCE)


of complex-(15)-sensitized solar cell shows a plateau between 500-600 nm reaching
83% at 570 nm. From the overlap integral of this curve with the AM-1.5 spectral
solar photon flux one derives a / sc of 17.1 mA/cm2. The open circuit voltage (Voc) was
802 mV and the fill factor (FF) 0.74 ± 0.03. Using these values one derives an overall
efficiency η of 10.08% [3.54].

3.3 ORGANIC SENSITIZERS

3.3.1 High efficiency organic sensitizers


Over the last 20 years, ruthenium complexes with remarkable efficiency have been
designed. However, there are several challenges that limit their development for large-
scale applications. Therefore several groups have developed metal-free organic sensi­
tizers for which they have obtained comparable efficiencies to the ruthenium sensitizers.
In particular, the high extinction coefficients of organic sensitizers are attractive for
use in solid-state dye-sensitized solar cells. A very efficient courmarine dye (16) for
Molecular engineering of sensitizers 103

Fig. 3.20 Computed isosurface plots of selected Kohn-Sham molecular orbitals (from HOMO-2
to LUMO + 2) of complex (15) which are mainly involved in the optically active electronic
transitions 1, 2 and 3 (black-bars in Figure 3.19).

DSCs showed a solar-to-electric power conversion yields approaching 7.7% in full


sunlight [3.55]. The bridging thiophene units are used to provide conjugation between
the donor and the anchoring groups, as well as to increase the molar extinction coef­
ficient of the dye. In order to extend optical response, the electron acceptor CN group
was introduced, which decreases the gap between HOMO and LUMO, thus leading
to the shift of the lowest energy absorption maximum from 511 to 552 nm [3.56].
Hence, dye-(17)-sensitized solar cells generated 25% higher / s c , due to its enhanced
light-harvesting properties. The molar extinction coefficients of both dye (16) and
(17) in ethanol were determined to be 64,300 (at 511 nm) and 97,400 M _ 1 cm - 1 (at
552 nm), respectively, which is about 5 and 7 times greater than that of complex (2).
An indoline dye (18) with an extinction coefficient of 68,700 M _ 1 cm - 1 at 526 nm
reaches 8% conversion efficiency when employed with a volatile organic electrolyte
[3.57]. After meticulous optimization, the dye-(18)-sensitized solar cell with a vola­
tile organic electrolyte showed an impressive response over a wide spectral range in
IPCE, where the values exceeded 80% between 410 and 670 nm, resulting in a high
/ s c value (approx. 20 mA cm - 2 ) and a conversion efficiency of 9.03%. Controlling
molecular aggregation on the semiconductor is one way to improve the efficiency
104 Dye-Sensitized Solar Cells

of DSCs because aggregation may lead to intermolecular quenching or molecules


residing in the system that are not functionally attached to the Ti0 2 surface, thereby
acting as filters [3.58]. To decrease /-aggregation of dye (18), a new indoline dye
(19) was designed by introducing an n-octyl substitute onto the rhodanine ring of dye
(18) [3.59]. Comparisons between dye (18) and (19) in Table 3.21 show that n-octyl
substitution increased the Voc regardless of the presence of co-adsorbent, in this case
chenodeoxylcholic acid (CDC A). With the combination of CDC A and the n-octyl
chain, dye (19) improves the Voc up to 0.710 V, which is 10.2% higher than that of dye
(18) with CDCA. As mentioned above, alkyl substitution of dyes has improved the Voc
due to the blocking effect of the charge recombination between oxidized species, i.e.
I3~ and electrons injected in the nanocrystalline-Ti02 electrodes. With supported high
Voc (0.717 V) and / sc (18.56 mA/cm2), 9.52% as best efficiency was achieved with
dye-(19)-sensitized solar cells [3.59].

Koumura et al. observed an effect of thiophene alkyl-chain length in the


^-conjugation bridge of organic dyes (20-21) [3.60]. An absorption maximum of
Molecular engineering of sensitizers 105

Table 3.21 Current density vs. voltage characteristics for DSCs with dye (18) and (19)
as sensitizers with and without CDC A under AM-1.5 simulated-sunlight (100 mW/cm2)
illumination.
Sensitizer / sc [mA/cm2] Ko [mV] Fill factor Efficiency [%]

Without CDCA
Complex (18) 19.08 + 0.26 638 ± 50 0.682 ± 0.06 8.26 ± 0.09
Complex (19) 18.99 + 0.19 656 + 110 0.678 ± 0.09 8.43 + 0.16
With CDCA
Complex (18) 19.86 + 0.10 644+130 0.694 ± 0.06 8.85 + 0.18
Complex (19) 18.68 + 0.08 710 + 70 0.707 ± 0.00 9.40 + 0.12

complex (21) was observed at 473 nm, which is red-shifted compared to the complex
(20) due to increased ^-conjugation. The IPCE for a DSSC based on complex (21)
showed the onset wavelength 800 nm, and the efficiency was 7.7% under AM-1.5 G
irradiation (100 mW/m2). The Voc of complex-(20) and complex-(21)-sensitized solar
cells were 80 mV higher than that for which no n-hexyl groups are substituted at
the oligothiophene moieties, due to longer electron lifetimes caused by alkyl chains,
which are effectively reducing molecular aggregation [3.60].

The tailored dimethylfluoreneaniline moieties in the complex-(22) and complex-


(23)-sensitizers ensure greater resistance to degradation under exposure to light and
high temperatures, compared to simple arylamines. The bridging thiophene units are
used to provide conjugation between the donor and the anchoring groups. The HOMO
of both sensitizers, in vacuum and in solution, is delocalized across the bis-dimethyl-
fluoreneaniline ligand, with maximum components arising from the nitrogen lone-
pair and the ;r framework of the surrounding ligands, see Figure 3.22. The LUMO is
in both cases a π* orbital, delocalized across the thiophene and cyanoacrylic groups,
106 Dye-Sensitized Solar Cells

Fig. 3.22 Isodensity surface plots of the HOMO and LUMO of dye (22). Similar orbitals are
calculated for dye (23).

Fig. 3.23 IPCE obtained with a nanocrystalline Ti02 film supported on FTO conducting glass
and derivatized with monolayer of dye (22) (dashed line) and (23) (solid line).

with sizable components from the cyano- and carboxylic moieties [3.61]. The IPCE
data of both sensitizers, plotted as a function of excitation wavelength, exhibit a strik­
ingly high plateau at 91%; and the IPCE of the dye-(23) solar cell is red shifted by
about 30 nm compared to the dye-(22) solar cell as a result of extended ^-conjuga­
tion (see Figure 3.23). Under standard global AM-1.5 solar conditions, the dye-(23)-
sensitized cell gave a short circuit photocurrent density (/sc) of 14.0 ± 0.20 mA/cm2,
an open circuit voltage of 753 ± 10 mV and a fill factor of 0.77 ± 0.02, corresponding
to an overall conversion efficiency η of 8.01% which is higher than the dye-(22)-
sensitized cell. The lower efficiency of dye (22) (-11%) compared to the dye-(23)
sensitizer demonstrates the benfitial influence of thiophene units on the photocur­
rent, due to the enhanced the red response. With its dithienothiophene unit fused with
bisfluorenylaniline, dye (24) showed comparable efficiency, 8.0%, and good stability
under light soaking conditions because the fused dithienothiophene unit has a low free
Molecular engineering of sensitizers 107

energy of solvation in the high polarity electrolytes normally used by DSCs [3.62].
With a solvent-free ionic-liquid electrolyte, the efficiency of the dye-(24)-sensitized
solar cell only changed from 6.7 to 6.1%, due to a 47-mV drop of Voc with no change
of / sc under light soaking conditions at 60°C during a period lasting 1000 h.

Organic dyes for DSCs based on diphenylaniline as the electron donor were
designed and synthesized [3.38, 3.40, 3.63-3.66]. The diphenylaniline moiety is
non-planar and would suppress aggregation through disturbance of the π-π stack­
ing [3.65]. The dyes (25-26) show the systematic effect of substituted donor groups
on controlling optical and photovoltaic properties [3.38]. The absorption spectra of
the dye (25) displays a strong visible band at 441 nm (ε= 33,000 M -1 cm-1) due to
the π-π* transitions of the conjugated molecule. However, the dye (26) (462 nm,
£= 33,000 M_1 cm-1), which contains methoxy groups, shows absorption maxima
red shifted compared to the dye (25) (see Fig. 3.24). This change in optical response
of dyes (25-26) are revealed in IPCE data, where all sensitizers plotted as a function
of excitation wavelength exhibit a high plateau at 85%. The IPCE spectra of dye (26),
which contains methoxy groups, exhibit a 30-nm red shift compared to dye (25),
108 Dye-Sensitized Solar Cells

Fig. 3.24 Normalized absorption spectra of dye (25) (black line) and (26) (red line) measured
in ethanol.

consistent with the absorption spectra as shown in Figure 3.25. The dye (27), for
which hexyloxyl chains are incorporated at the diphenylaniline donor ligand, showed
an overall conversion efficiency of 7.25% with volatile electrolyte and 6% with ionic
liquid, and remained at 90% after 1000 hours [3.40].
Molecular engineering of sensitizers 109

Fig. 3.25 The IPCE obtained with a nanocrystalHne Ti02 film supported on FTO conducting
glass and derivatized with monolayer of dye (25) (black square) and (26) (red circle).

3.3.2 Near-IR absorbing sensitizers


The main drawback of the ruthenium-based sensitizers and the organic sensitizers intro­
duced above is the lack of absorption in the red region of the visible spectrum. In this
section, sensitizers having intense absorption in the near-infrared regions are introduced.
Phthalocyanines (Pcs) exhibit absorption maxima near 700 nm, with very high extinc­
tion coefficients (where the maximum of the solar photon flux occurs), that make them
especially suitable for integration in light-energy-conversion systems [3.67]. However
Pcs have shown unimpressive power conversion efficiencies [3.68-3.72], and the low
efficiency of cells incorporating phthalocyanines appears to be due to aggregation and
lack of directionality in the excited state [3.73]. One of the essential properties required
for the light-harvesting system of a molecular/semiconductor junction is that the sen-
sitizer be in the excited state. This directionality should be created by engineering the
sensitizer at the molecular level to provide efficient electron transfer from the excited
dye to the Ti0 2 conduction band, via good electronic coupling between the lowest unoc­
cupied molecular orbital (LUMO) of the dye and the Ti 3d orbital. In the dye (28), three
tert-butyl and two carboxylic acid groups act to push and pull, respectively. The func­
tion of the two carboxylic acid groups is to graft the sensitizer on the semiconductor sur­
face, and to provide intimate electronic coupling between its excited state wave-function
and the conduction-band manifold of the semiconductor. The purpose of the three tert-
butyl groups is to enhance solubility, minimize aggregation, and tune the LUMO level
of the Pc so that it provides directionality in the excited state. Therefore, the incident
monochromatic photon-to-current conversion efficiency (IPCE), plotted as a function
of excitation wavelength, reaches 75% and a conversion efficiency that exceeds 3% at 1
sun [3.74]. The dye (29), which contains anchoring groups directly attached on phthalo-
cyanine, has demonstrated efficient electron injection to Ti0 2 [3.75]. Figure 3.26 shows
time-correlated single-photon counting (TCSPC) to estimate the electron injection
yield of dye (29) on the mesoporous semiconductor. The excited lifetime of dye (29),
when adsorbed onto mesoporous metal-oxide films, is short (Ti = 1.1 ns (81.4%) and
110 Dye-Sensitized Solar Cells

Fig. 3.26 Time-correlated single-photon-counting measurements for dye (29) in EtOH 10 6 M


solution with 2 x 10~4 M chenodeoxycholic acid and A1203 film. λ6Χ0 = 635 nm, Xemi = 695 nm.

τ2 = 2.8 ns (18.6%)), when compared to the solution sample (τϊ = 4 n s (40.5%) and
τ2 = 2.8 ns (59.5%)). An injection yield higher than 80% was measured by compar­
ing the signal intensity between dye (29)/Al 2 0 3 and dye (29)/Ti0 2 while keeping the
acquisition time constant. The DSC based on dye (29) shows a record efficiency under
solar-simulated light irradiation (100 mW/cm 2 1.5-AM Global) of 3.52% with 80% of
IPCE at 690 nm after optimization (see Fig. 3.27) [3.75, 3.76].

Another well known class of compounds for intense absorption in the red/near-
IR regions is the squaraines [3.77-3.80]. An asymmetric squaraine sensitizer, dye
Molecular engineering of sensitizers 111

Fig. 3.27 The IPCE spectrum (insert) and I/V curve of dye (29)/Ti02 DSC with an active area
of 0.2 cm2.

Fig. 3.28 Absorption spectra of dye (30) in ethanol solution (black solid line) and onto 4-/mi
Ti02film(red dotted line).

(30), shows a strong absorption maximum at 636 nm with high molar extinction coef­
ficient (ε= 158,500 M_1 cm-1), which is due to π-π* charge transfer (CT) transitions
(see Fig. 3.28) [3.81]. Inspection of the electronic structure of dye (30) revealed that
the HOMO is delocalized throughout the dye, while the HOMO-1, 0.23 eV below the
HOMO, is entirely localized within the squaraine core, with both orbitals belonging
to the ;r framework of the dye, as seen in Figure 3.29. The LUMO, 1.32 eV above the
HOMO, is a /r* orbital delocalized throughout the dye, with sizable contributions aris­
ing from the carboxylic group, while the LUMO + 1, calculated to be 1.13 eV above
the LUMO, is entirely localized in the dye portion bearing the carboxylic substituent.
The intense 596 nm transition is mainly composed of HOMO-LUMO and HOMO-
LUMO + 1 excitations, the former being the dominant contribution. This transition
essentially gives rise to a charge flow from the squaraine core to the outer molecular
region, where the carboxylic group is located, thus producing an excited state which
is potentially strongly coupled to the semiconductor surface, to which the carboxylic
group is linked. Figure 3.30 shows the IPCE and the photovoltaic performance of the
112 Dye-Sensitized Solar Cells

Fig. 3.29 Upper panel: isodensity plot of selected molecular orbitals of the squaraine dye (30).
Lower panel: charge density difference between the excited (596 nm) and ground state; the
yellow (white) color indicates an increase (decrease) of charge density in a given molecular
region.

Fig. 3.30 Photocurrent-action spectrum (insert) and current-voltage characteristics of dye (30)
obtained with a nanocrystalline Ti0 2 film, supported onto a conducting glass sheet and deriva-
tized with monolayer of squaraine in presence of chenodeoxycholic acid.

dye-(30)-sensitized solar cell. The IPCE of the dye-(30)-sensitized solar cell, plotted
as a function of excitation wavelength shows 85% and under standard global AM-1.5
solar condition, gave a short-circuit photocurrent density of 10.50 ± 0.20 mA/cm 2 , an
open-circuit voltage of 603 ± 30 mV and a fill factor of 0.71 ± 0.03, corresponding to
an overall conversion efficiency, 77, of 4.50% [3.81]. UV-Vis spectra of red and/or near
IR absorbing dyes exhibit an optical window in part of the visible region which, with
Molecular engineering of sensitizers 113

Fig. 3.31 Photocurrent-action spectrum of dye (30) (red line), (23) (blue line), and co-
sensitization (30)/(23) (black line) obtained with a nanocrystalline Ti02film,supported onto a
conducting glass sheet.

the appropriate dye combination, can be used to achieve a panchromatic sensitization,


the so called co-sensitization [3.82-3.85]. Dye (30) (or dye (29) [3.75]), together with
dye (23), showed successful co-sensitization [3.86]. The co-sensitization broadens
the photocurrent-action spectrum, covering the entire visible domain with peak effi­
ciencies of 86 and 76% at 530 and 660 nm, respectively, corresponding to the highest
absorption of dye (23) and (30) (see Fig. 3.31).

3.4 REFERENCES

[3.1] International Energy Outlook 2008, Energy Information Administration, 2008, http://www.eia.doe.
go v/oiaf/ieo/index. html.
[3.2] Grätzel M., Nature 2001, 474, 338.
[3.3] Nazeeruddin, M. K., Grätzel, M., Molecular and Supramolecular Photochemistry, Vol. 10, Marcel
Dekker, New York, 2003.
[3.4] Nazeeruddin, M. K., Special issue: Michael Graetzel Festschrift, a tribute for his 60th Birthday:
Dye Sensitized Solar Cells, Vol. 248, Elsevier, Amsterdam, 2004.
[3.5] Anderson, S., Constable, E. C., Dareedwards, M. P., Goodenough, J. B., Hamnett, A., Seddon, K.
R., Wright, R. D., Nature 1979, 280, 571.
[3.6] O'Regan, B., Grätzel, M., Nature 1991, 353,131.
[3.7] Grätzel, M., Handbook of nanostructured materials and nanotechnology, Vol. 3, Academic Press,
2000.
114 Dye-Sensitized Solar Cells

[3.8] Nazeeruddin, M. K., Grätzel, M., Encyclopedia of Electrochemistry: Semiconductor Electrodes and
Photoelectrochemistry, Vol. 6, Wieley-VCH, Germany, 2002.
[3.9] Green, M. A., Emery, K., Hisikawa, Y., Warta, W., Prog. Photovolt: Res. Appl. 2007, 75, 425.
[3.10] Green, M. A., Emery, K., Hishikawa, Y., Warta, W., Prog. Photovolt: Res. Appl. 2009,17, 85.
[3.11] Haught, A. E, Journal of Solar Energy Engineering-Transactions of the Asme 1984,106,3.
[3.12] De Vos, A., Endoreversible Thermodynamics of Solar Energy Conversion, Vol. chapter 6, Science
Publishers, Oxford, 1992.
[3.13] Nazeeruddin, M. K., Zakeeruddin, S. M., Humphry-Baker, R., Jirousek, M., Liska, P.,
Vlachopoulos, N., Shklover, V., Fischer, C. H., Grätzel, M., Inorg. Chem. 1999, 38, 6298.
[3.14] Hara, K., Sugihara, H., Tachibana, Y, Islam, A., Yanagida, M., Sayama, K., Arakawa, H.,
Fujihashi, G., Horiguchi, T., Kinoshita, T., Langmuir 2001,17, 5992.
[3.15] Yan, S. G., Hupp, J. T., /. Phys. Chem. 1996,100, 6867.
[3.16] Tachibana, Y, Haque, S. A., Mercer, I. P., Moser, J. E., Klug, D. R., Durrant, J. R., /. Phys. Chem.
B 2001,105, 7424.
[3.17] Nazeeruddin, M. K., Humphry-Baker, R., Liska, P., Grätzel, M., /. Phys. Chem. B 2003, 107,
8981.
[3.18] Péchy, P., Rotzinger, F. P., Nazeeruddin, M. K., Kohle, O., Zakeeruddin, S. M., Humphrybaker, R.,
Grätzel, M., /. Chem. Soc. Chem. Commun. 1995, 65.
[3.19] Heimer, T. A., Darcangelis, S. T., Farzad, F., Stipkala, J. M., Meyer, G. J., Inorg. Chem. 1996, 35,
5319.
[3.20] Campus, F, Bonhote, P., Grätzel, M., Heinen, S., Walder, L., Sol. Energy Mater. Sol. Cells 1999,
56,281.
[3.21] Rice, C. R., Ward, M. D., Nazeeruddin, M. K., Grätzel, M. New Journal of Chemistry 2000, 24, 651.
[3.22] Wang, Z. S., Li, F. Y, Huang, C. H., /. Phys. Chem. B 2001,105, 9210.
[3.23] Maier, S. A., Kik, P. G., Atwater, H. A., Meltzer, S., Harel, E., Koel, B. E., Requicha, A. A. G., Nat.
Mater. 2003, 2, 229.
[3.24] Wang, P., Zakeeruddin, S. M., Moser, J. E., Nazeeruddin, M. K., Sekiguchi, T., Grätzel, M., Nat.
Mater. 2003, 2, 402.
[3.25] Nazeeruddin, M. K., Kay, A., Rodicio, I., Humphrybaker, R., Muller, E., Liska, P., Vlachopoulos,
N., Grätzel, M., /. Am. Chem. Soc. 1993, 775, 6382.
[3.26] Huang, S. Y, Schlichthörl, G., Nozik, A. J., Grätzel, M., Frank, A. J., /. Phys. Chem. B 1997, 707,
2576.
[3.27] Kay, A., Grätzel, M., /. Phys. Chem. B 1993, 97, 6272.
[3.28] Schlichthörl, G., Huang, S. Y, Sprague, J., Frank, A. J., /. Phys. Chem. B 1997, 707, 8141.
[3.29] Wang, P., Zakeeruddin, S. M., Comte, P., Charvet, R., Humphry-Baker, R., Grätzel, M., /. Phys.
Chem. B 2003, 707, 14336.
[3.30] Neale, N. R., Kopidakis, N., van de Lagemaat, J., Grätzel, M., Frank, A. J., /. Phys. Chem. B 2005,
709,23183.
[3.31] Zhang, Z. P., Zakeeruddin, S. M., O'Regan, B. C , Humphry-Baker, R., Grätzel, M. /. Phys. Chem.
5 2005,709,21818.
[3.32] Yum, J. H., Moon, S. J., Humphry-Baker, R., Walter, P., Geiger, T., Nuesch, F., Grätzel, M.,
Nazeeruddin, M. D. K., Nanotechnology 2008, 79.
[3.33] Tennakone, K., Bandara, J., Bandaranayake, P. K. M., Kumara, G. R. A., Konno, A., Jpn. J. Appl.
Phys., Part 2 2001, 40, L732.
[3.34] Kay, A., Grätzel, M., Chem. Mater. 2002,14, 2930.
[3.35] Palomares, E., Clifford, J. N., Haque, S. A., Lutz, T., Durrant, J. R., /. Am. Chem. Soc. 2003, 725,
475.
[3.36] Gregg, B. A., Pichot, F, Ferrere, S., Fields, C. L. /. Phys. Chem. B 2001, 705, 1422.
[3.37] Kroeze, J. E., Hirata, N., Koops, S., Nazeeruddin, M. K., Schmidt-Mende, L., Grätzel, M.,
Durrant, J. R., /. Am. Chem. Soc. 2006,128, 16376.
[3.38] Hagberg, D. P., Yum, J. H., Lee, H., De Angelis, F, Marinado, T., Karlsson, K. M., Humphry-Baker, R.,
Sun, L. C, Hagfeldt, A., Grätzel, M., Nazeeruddin, M. K., /. Am. Chem. Soc. 2008,130, 6259.
[3.39] Xu, M. F , Li, R. Z., Pootrakulchote, N., Shi, D., Guo, J., Yi, Z. H., Zakeeruddin, S. M., Grätzel, M.,
Wang, P., /. Phys. Chem. C2008, 772, 19770.
Molecular engineering of sensitizers 115

[3.40] Yum, J. H., Hagberg, D. P., Moon, S. J., Karlsson, K. M., Marinado, T., Sun, L. C., Hagfeldt, A.,
Nazeeruddin, M. K., Grätzel, M.,Angew. Chem., Int. Ed. 2009, 48, 1576.
[3.41] Zakeeruddin, S. M., Nazeeruddin, M. K., Humphry-Baker, R., Péchy, P., Quagliotto, P., Barolo, C.,
Viscardi, G., Grätzel, M., Langmuir 2002,18, 952.
[3.42] Jang, S. R., Yum, J. H., Klein, C., Kim, K. J., Wagner, P., Officer, D., Grätzel, M., Nazeeruddin, M. K.,
/. Phys. Chem. C2009,113, 1998.
[3.43] Nazeeruddin, M. K., Wang, Q., Cevey, L., Aranyos, V., Liska, P., Figgemeier, E., Klein, C., Hirata,
N., Koops, S., Haque, S. A., Durrant, J. R., Hagfeldt, A., Lever, A. B. P., Grätzel, M., Inorg. Chem.
2006, 45, 787.
[3.44] Nazeeruddin, M. K., Bessho, T., Cevey, L., Ito, S., Klein, C., De Angelis, F., Fantacci, S., Comte,
P., Liska, P., Imai, H., Graetzel, M., /. Photochem. Photobiol, A 2007,185, 331.
[3.45] Anderson, P. A., Strouse, G. F , Treadway, J. A., Keene, F R., Meyer, T. J., Inorg. Chem. 1994, 33,
3863.
[3.46] Alebbi, M., Bignozzi, C. A., Heimer, T. A., Hasselmann, G. M., Meyer, G. J., /. Phys. Chem. B
1998,102, 7577.
[3.47] Treadway, J. A., Moss, J. A., Meyer, T. J., Inorg. Chem. 1999, 38, 4386.
[3.48] Nazeeruddin, M. K., Péchy, P., Grätzel, M., Chem. Commun. 1997, 1705.
[3.49] Grätzel, M., /. Photochem. Photobiol., A 2004,168, 235.
[3.50] Nazeeruddin, M. K., Péchy, P., Renouard, T., Zakeeruddin, S. M., Humphry-Baker, R., Comte, P.,
Liska, P., Cevey, L., Costa, E., Shklover, V., Spiccia, L., Deacon, G. B., Bignozzi, C. A., Grätzel,
M., /. Am. Chem. Soc. 2001,123, 1613.
[3.51] Chiba, Y, Islam, A., Watanabe, Y, Komiya, R., Koide, N., Han, L. Y, Jpn. J. Appl. Phys., 2 2006,
45, L638.
[3.52] Nazeeruddin, M. K., De Angelis, F , Fantacci, S., Selloni, A., Viscardi, G., Liska, P., Ito, S.,
Bessho, T., Grätzel, M., /. Am. Chem. Soc. 2005, 727, 16835.
[3.53] Wadman, S. H., Kroon, J. M., Bakker, K., Lutz, M., Spek, A. L., van Klink, G. P. M., van Koten, G.,
Chem. Comm. 2007, 1907.
[3.54] Bessho, T., Yoneda, E., -Yum, J.-H., Guglielmi, M., Tavernelli, I., Imai, H., Rothlisberger, U.,
Nazeeruddin, M. K., Grätzel, M., /. Am. Chem. Soc. 2009,131, 5930.
[3.55] Hara, K., Kurashige, M., Dan-oh, Y, Kasada, C , Shinpo, A., Suga, S., Sayama, K., Arakawa, H.,
N. J. Chem. 2003, 27, 783.
[3.56] Wang, Z. S., Cui, Y, Hara, K., Dan-Oh, Y, Kasada, C , Shinpo, A., Adv. Mater. 2007,19, 1138.
[3.57] Horiuchi, T., Miura, H., Sumioka, K., Uchida, S., /. Am. Chem. Soc. 2004,126, 12218.
[3.58] Khazraji, A. C , Hotchandani, S., Das, S., Kamat, P. V., /. Phys. Chem. B 1999,103, 4693.
[3.59] Ito, S., Miura, H., Uchida, S., Takata, M., Sumioka, K., Liska, P., Comte, P., Pechy, P., Graetzel, M.,
Chem. Commun. 2008, 5194.
[3.60] Koumura, N., Wang, Z. S., Mori, S., Miyashita, M., Suzuki, E., Hara, K., /. Am. Chem. Soc. 2006,
128, 14256.
[3.61] Kim, S., Lee, J. K., Kang, S. O., Ko, J., Yum, J. H., Fantacci, S., De Angelis, F, Di Censo, D.,
Nazeeruddin, M. K., Grätzel, M., /. Am. Chem. Soc. 2006,128, 16701.
[3.62] Qin, H., Wenger, S., Xu, M., Gao, F, Jing, X., Wang, P., Zakeeruddin, S. M., Grätzel, M., /. Am.
Chem. Soc. 2008,130, 9202.
[3.63] Kitamura, T., Ikeda, M., Shigaki, K., Inoue, T., Anderson, N. A., Ai, X., Lian, T. Q., Yanagida, S.,
Chem. Mater. 2004,16, 1806.
[3.64] Velusamy, M., Thomas, K. R. J., Lin, J. T., Hsu, Y. C , Ho, K. C , Org. Lett. 2005, 7, 1899.
[3.65] Hagberg, D. P., Edvinsson, T., Marinado, T., Boschloo, G., Hagfeldt, A., Sun, L. C , Chem. Commun.
2006, 2245.
[3.66] Hwang, S., Lee, J. H., Park, C , Lee, H., Kim, C , Park, C , Lee, M. H., Lee, W., Park, J., Kim, K.,
Park, N. G., Kim, C , Chem. Commun. 2007, 4887.
[3.67] Leznoff, C. C , Lever, A. B. P., in Phthalocyanines, Vol. 3, VCH publishers, Inc., New York, 1993.
[3.68] Giraudeu, A., -Fan, F-R. F , Bard, J., /. Am. Chem. Soc. 1980,102, 5137.
[3.69] Wöhrle, D., Meissner, D., Adv. Mater. 1991, 3, 129.
[3.70] He, J., Hagfeldt, A., -Lindquist, S.-E., Grennberg, H., Korodi, F, Sun, L., Akermark, B., Langmuir
2001,17, 2743.
116 Dye-Sensitized Solar Cells

[3.71] He, J., Benko, G., Korodi, F., Polivka, T., Lomoth, R., Âkermark, B., Sun, L., Hagfeldt, A.,
Sundstrom, V., /. Am. Chem. Soc. 2002, 724, 4922.
[3.72] Palomares, E., Martinex-Diaz, M. V., Haque, S. A., Torres, T., Durrant, J. R., Chem. Chommun.
2004,2112.
[3.73] Nazeeruddin, M. K., Humphry-Baker, R., Grätzel, M., Wöhrle, D., Schnurpfeil, G., Schneider, G.,
Hirth, A., Trombach, N., /. Porphyrins Phthalocyanines 1999, 3, 230.
[3.74] Reddy, P. Y., Giribabu, L., Lyness, C , Snaith, H. J., Vijaykumar, C , Chandrasekharam, M.,
Lakshmikantam, M., Yum, J. H., Kalyanasundaram, K., Grätzel, M., Nazeeruddin, M. K., Angew.
Chem., Int. Ed. 2007, 46, 373.
[3.75] Cid, J. J., Yum, J. H., Jang, S. R., Nazeeruddin, M. K., Ferrero, E. M., Palomares, E., Ko, J.,
Grätzel, M., Torres, T., Angew. Chem., Int. Ed. 2007, 46, 8358.
[3.76] Yum, J. H., Jang, S. R., Humphry-Baker, R., Grätzel, M., Cid, J. J., Torres, T., Nazeeruddin, M. K.,
Langmuir 2008, 24, 5636.
[3.77] Alex, S., Santhosh, IL, Das, S., /. Photochem. Photobiol, A 2005, 772, 63.
[3.78] Li, C , Wang, W., Wang, X. S., Zhang, B. W., Cao, Y, Chem. Lett. 2005, 34, 554.
[3.79] Otsuka, A., Funabiki, K., Sugiyama, N., Yoshida, T., Chem. Lett. 2006, 35, 666.
[3.80] Burke, A., Schmidt-Mende, L., Ito, S., Grätzel, M., Chem. Commun. 2007, 234.
[3.81] Yum, J. H., Walter, P., Huber, S., Rentsch, D., Geiger, T., Nuesch, F., De Angelis, F., Grätzel, M.,
Nazeeruddin, M. K., /. Am. Chem. Soc. 2007, 729, 10320.
[3.82] Ehret, A., Stuhl, L., Spitler, M. T., /. Phys. Chem. B 2001,105, 9960.
[3.83] Sayama, K., Tsukagoshi, S., Mori, T., Hara, K., Ohga, Y, Shinpou, A., Abe, Y, Suga, S., Arakawa, H.
Sol. Energy Mater. Sol. Cells 2003, 80, 47.
[3.84] Chen, Y. S., Zeng, Z. H., Li, C , Wang, W. B., Wang, X. S., Zhang, B. W. New J. Chem. 2005,
29, 773.
[3.85] Perera, V. P. S., Pitigala, P. K. D. D. P., Senevirathne, M. K. I., Tennakone, K., Sol. Energy Mater.
Sol. Cells 2005, 85, 91.
[3.86] Yum, J. H., Jang, S. R., Walter, P., Geiger, T., Nuesch, F , Kim, S., Ko, J., Grätzel, M.,
Nazeeruddin, M. K., Chem. Commun. 2007, 4680.
CHAPTER 4

OPTIMIZATION OF REDOX MEDIATORS AND


ELECTROLYTES
Ke-jian Jiang and Shozo Yanagida*

4.1 INTRODUCTION

Dye-sensitized solar cells (DSCs) or "Grätzel cells" have attracted widespread aca­
demic and commercial interest during the last 18 years due to their potential for low-
cost solar energy conversion [4.1-4.10]. With the optimization of the key materials
and device engineering, high conversion efficiencies of 10-11 % have been achieved
with organic solvent-based electrolytes under standard air mass 1.5 (AM1.5G) glo­
bal sunlight [4.2-4.4]. Moreover, conversion efficiencies of around 9 % have been
reported for ion liquid-based electrolytes with long-term device stability [4.4].
In DSCs, a typical iodide/triiodide redox couple is usually employed as an effi­
cient electron mediator in an electrolyte. It is sandwiched between a working electrode
and a counter electrode, as shown in Figure 4.1. The function of the redox couple is
to reduce the dye cation, following electron injection, and carry the charge back and
forth between the two electrodes. The redox couple is thus of crucial importance
for a stable operation and high performance of the device. Generally, an electrolyte
for DSCs consists of a redox couple, and various additives, such as lithium salt and
terf-butylpyridine, in solvents. According to its physical states, an electrolyte can be
roughly divided into one of three types: liquid electrolyte, quasi-solid state electro-
lyte, and solid-state electrolyte. Liquid electrolytes can be further divided into organic
solvent-based electrolytes and ionic liquid-based electrolytes, according to the sol­
vent used.
Organic solvent-based electrolytes (coded as OLEs) have been widely used and
investigated in DSCs with high light-to-electricity conversion efficiencies since vola­
tile organic solvents, such as acetonitrile, can easily dissolve a wide range of organic
and inorganic compounds. The electrolyte can thus fully pervade in the mesoporous
dyed Ti0 2 film and efficiently carry out a series of physicochemical processes. At
full sunlight (100 mW cm-2), a certified efficiency of 11.1 % was reported in a DSC
118 Dye-Sensitized Solar Cells

Fig. 4.1 A schematic structure of a dye-sensitized solar cell.

with an OLE consisting of 0.6 M dimethylpropyl imidazolium iodide, 0.1 M lithium


iodide, 0.05 M iodine, and 0.5 M terf-butylpyridine in acetonitrile [4.3].
In consideration of long-term stability and practical application of DSCs, non-vol­
atile ionic liquid-based electrolytes (coded as ILEs), quasi-solid state electrolytes, and
solid-state electrolytes have been developed as desirable alternatives to OLEs. In DSCs
with an ILE consisting of l,3-dimethylimidazolium/l-ethyl-3-methylimidazolium tetra-
cyanoborate/12/N-butylbenzoimidazole/guanidium thiocyanate, molar ratio: 12/12/16/1.
67/3.33/0.67, efficiencies of 8.5-9.5 % were recently reported, with high stability, under
full sunlight soaking at 60 °C for 1000 firs [4.4]. In further consideration of manufactur­
ing procedures and practical applications to match various architectural constructions,
quasi- or solid-state electrolytes may be desirable, in which the liquid electrolyte is
gelatinized by adding gelling agents (inorganic nanoparticles, low molecular weight
and polymer gelators). Comparable photovoltaic parameters to those for DSCs without
gelators in ionic liquid electrolyte have been reported.
In this chapter, we focus on organic solvent-based electrolytes, non-volatile
ionic liquid-based electrolytes, and quasi-solid state electrolytes, where an I~/T redox
couple is employed. The iodine-free solid electrolytes, including organic and inor­
ganic semiconductors [4.5-4.7], are discussed elsewhere. The charge transfer proc­
esses in the DSCs are first presented, after which a variety of components and their
roles in the electrolyte are investigated. Recent progress with regard to DSCs with
either ionic-liquid or quasi-solid state electrolytes is also discussed.

4.2 CHARGE TRANSFER PROCESSES IN DSCs

Charge transfer processes have been extensively investigated in dye-sensitized


nanocrystalline Ti0 2 solar cells with electrolytes typically containing an iodide/
Optimization of redox mediators and electrolytes 119

triiodide redox couple, as shown in Figure 4.2 [4.11-4.18]. Under light illumination,
the photoexcited electrons are injected from the dye sensitizers (D*) into the con­
duction band of the nanocrystalline semiconductor [Eq. 4.1], and the electrons dif­
fuse through the semiconductor and reach the external circuit. The resulting dye
cations (D + ) are reduced by iodide anions, while I~ ions are oxidized in the form
of I~ [Eq. 4.2]. The I~ ions diffuse toward a platinized counter electrode, where
they become reduced by accepting electrons from the external circuit. The charge
separation in a nanocrystalline semiconductor is thought not to depend on a built-in
electrical field, i.e., a Shottky barrier, but is mainly determined by electron kinet­
ics at the semiconductor/dye/electrolyte interface. In an efficient DSC, the rate of
electron injection, Km], must be higher than the decay rate of the dye-excited state to
ground, and the rate of recombination of the dye cation with electrons injected into
the semiconductor [Eq. 4.3], Kh2, must be lower than the rate of reduction of the dye
cation by iodide, Kn. In DSCs with electrolytes employing an I / I " redox couple,
the electron injection process occurs on femtosecond-picosecond time scales with
a high quantum yield for most sensitizers. Moreover, the oxidized dye molecules
can be rapidly reduced by I~ ions present at the high concentrations (typically 0.5 M
in the electrolyte) in OLEs. Thus, the back combination (Kh2) between the oxi­
dized dye and the injected electrons in the T i 0 2 conduction band can be negligible.

Fig. 4.2 A schematic description of a DSC, including the main processes. Optical excitation
of the dye gives rise to an ultrafast electron injection into Ti0 2 conduction band states (Xinj);
the oxidized dye cation is rereduced by the Γ/Ι" redox couple present in the electrolyte (Krr).
This process is followed by back electron transfer from electrons photoinjected into the Ti0 2
conduction band to the dye cation (Xbl), and the recombination reaction (Kb2) between injected
electrons and oxidized redox species in the electrolyte (I~ ions).
120 Dye-Sensitized Solar Cells

Due to low diffusion coefficients of electrons (Dn, < 10~4 cm2 s-1) on a Ti0 2 film
as compared to in a single crystal Ti0 2 , the back combination can occur through
electron transfer from Ti0 2 to I~ at the Ti0 2 particle/electrolyte interface [Eq. 4.4].
This recombination (Khi) is usually thought to be responsible for the dark current
and represents a main energy-loss channel in the photoelectrochemical process of
DSCs. In principle, back electron transfer from Sn0 2 of the glass-conducting layer
to I3 ions may occur due to the porous nature of the Ti0 2 film, but this process is
often ignored in consideration of the small area of naked Sn0 2 as compared to the
large Ti0 2 surface.

(4.1)

(4.2)

(4.3)

(4.4)

(4.5)

During the charge transfer processes, the oxidation of Γ takes place on the sur­
face of a dyed Ti0 2 film, and the oxidized species, I~, must migrate from the photoelec-
trode to the couterelectrode where I~ is reduced to iodide (Eq. 4.5). Simultaneously,
the reduced iodide must diffuse to the photoelectrde within the cell. In these proc­
esses, the iodide is depleted at the photoanode, and triiodide is depleted at the Pt
cathode, depending on the mass transport of the iodide and triiodide in the electrolyte.
In typical electrolytes, sufficient excess of Γ is employed, usually ten times greater
than the corresponding I~ concentration. In addition, the diffusion coefficient of Γ
is higher than that of I~. Thus, the current limitation in DSCs is mainly due to dif­
fusion of the triiodide ion. Generally, the conductivity of electrolyte solution can be
expressed by diffusional ion migration. In DSCs, however, the redox couple often
causes an electronic conduction process, and the electron exchange between the redox
couple plays a role in electronic conduction, especially in viscous ILEs with rather
packed polyiodide species due to the required high concentration of I~. As a result,
the Grotthus mechanism was put forward [4.19, 4.20], coupling electron hopping and
the chemical bonds of polyiodide exchange. This mechanism can be described by the
Dahms-Ruff equation [4.21]:

(4.6)

where Dapp is the physical diffusion coefficient, k&x is the electron exchange rate con­
stant, σ is the equilibrium center-to-center distance at the exchange reaction, and
c is the total concentration of the redox couple in the electrolyte. The second part
(k&xa2c/6) of Eq. 4.6 responds to the electron exchange diffusion coefficient (DGX), and
can be represented as follows [4.20]:
Optimization of redox mediators and electrolytes 121

4.3 ELECTROLYTE COMPONENTS AND THEIR


ROLES IN THE DSCs

4.3.1 Organic solvents


An organic solvent is used to dissolve redox mediators and various additives for the
preparation of OLEs, and to help charge transfer as media [4.22-4.25]. In DSCs, sol­
vents used in electrolytes can affect charge dissociation, separation and transfer, as
well as the photovoltaic performance [4.1-4.4].
In DSCs, the flatband potential (V^) of a metal oxide semiconductor electrode
(Ti02) is sensitive to the solvent used. For water and nonaqueous protic solvents, it
has been found that the flatband potential (V^) of Ti0 2 is significantly more positive
than for nonaqueous protic solvents such as acetonitrile due to proton adsorption-
desorption equilibrium in the protic solvents at the semiconductor-electrolyte solution
interface. This results in the chosen electrolytes being independent of the various cati­
ons, such as Li+ion, used in the electrolyte [4.1]. Solvent effects on the performance
of a DSC usually relate to the change of V^, and lead to the open-circuit photo-voltage
(VQC) and short-circuit photo-current (/sc) of a device becoming affected. Generally,
dipolar aprotic solvents are preferred as they offer a wider range of electrochemical
stability and greater solubility for redox mediators and the various additives used in
the electrolyte.
It has also been reported that the DSC performance decreased dramatically with
increasing irradiation time for the electrolyte with nonaqueous protic solvents such as
alcoholic solvents. A solvent with a strong donor ability, such as dimethyl sulfoxide
(DMSO), can negatively shift the conduction band level due to its basic property,
similar to that of the ter-butylpyridine (Y-BP) used as an additive in typical OLEs. The
Voc thus increases, but / sc decreases due to the diminution of the driving force of the
charge injection from the excited dye molecules to Ti0 2 . In OLEs with basic solvents
such as DMSO and DMF, dye molecules have been found to desorb from the Ti0 2
surface, giving rise to deterioration of the device performance. In consideration of the
above limitations, nitrile solvents, such as acetonitrile, 3-methoxypropionitrile and
valeronitrile, as well as esters such as ethylene carbonate (EC) and propylene carbon­
ate (PC), are therefore usually preferable as solvents for OLEs thanks to their favora­
ble electrochemical stability and appropriate polarity when employed together with
redox mediators, additives. It has been reported that optimized OLEs with a mixture
of solvents of tetrahydrofuran (THF, 20 vol %) and actetonitrile (80 vol %) containing
0.3 M tert-BP, provided a short-circuit photocurrent of 18.23 mA/cm2, an open-circuit
voltage of 0.73 V, a fill factor of 0.73, and an overall power conversion efficiency of
9.74 % at AM 1.5 sunlight [4.25].

4.3.2 Cations
Cations such as Li+, as counter ions of Γ and I" in the electrolytes, play an important
role in DSCs [4.26-4.34]. They can efficiently screen photo-injected electrons on the
Ti0 2 film, ensuring that electro-neutrality is obeyed through the Ti0 2 network. In such
122 Dye-Sensitized Solar Cells

cases, ion diffusion in an electrolyte is strongly correlated with electron transport in


the Ti0 2 film. The motion of electrons creates a charge imbalance, and the resulting
electric field drags the cations along with it. At the same time, negatively charged ions
are repelled from the electrons. Thus, the simultaneous motion of electrons and ions
yields a single diffusion coefficient: the ambipolar diffusion coefficient, £>Amb, which
is expressed as:

(4.7)

Here, n and Dn are the negative charge density and diffusion coefficient, respec­
tively, whereas p and Dp are their negative counterparts [4.27].
In a previous report [4.30], it was put forward that the nature and concentration
of cation species in the electrolyte had a strong influence on the efficiency of DSCs.
For instance, a high photocurrent was observed in the above-mentioned DSCs with
the Li+-containing electrolyte as opposed to the tetrapropyl ammonium-containing
electrolyte. The cations have been found to influence the open-circuit voltages, the
electron injection yield, and the rate of dye regeneration, and finally affect the photo­
voltaic efficiency of the device.
The cation effect on cell performance was systematically investigated for elec­
trolytes containing a series of alkaline metal ions: Lil, Nal, KI, Rbl, or Csl [4.30].
It was found that the short-circuit photocurrent (/sc) decreased, and the open-circuit
photo voltage (Voc) increased with the increase in cation radius from Li+ to Cs+ for
a wide range of light intensities. The dependence of / sc and Voc on the cation was
explained as due to the change in interfacial energetics of a dyed Ti0 2 film, and a
general conclusion was the following: upon light illumination of dyed nanoporous
Ti0 2 film, the injected electrons accumulated in the conduction band of Ti0 2 , and
the cations in electrolyte became adsorbed or intercalated in the Ti0 2 film for charge
compensation. The flat-band potential of Ti0 2 moved positively due to the absorp­
tion of cations, and the drop in potential depended on the size and charge density
of the employed cations. As compared to the cation of large size, the small-sized
cation could easily adsorb onto the Ti0 2 surface, which resulted in a large potential
drop. Due to the positive movement of the Ti0 2 flat band potential, the driving force
of the charge injection from the excited dye molecules to Ti0 2 increased, resulting
in a large injection efficiency and thus a high photocurrent. On the other hand, the
cation adsorption can affect the open-circuit voltage due to the fact that the voltage is
decided by the difference in photo-conduction band level and the redox potential of
the electrolyte. The cation effect strongly depends on the cation density in the electro­
lyte and the adsorption ability on the Ti0 2 film. A small size and high cation density
can be favorable to the adsorption. For instance, Li+ was found to intercalate into the
Ti0 2 lattice, and its effect on the short-circuit current and open circuit voltage was
more remarkable when its concentration in the electrolyte was high.
Cation effects on the electron transfer properties on a Ti0 2 film were inves­
tigated in the Ti02-electrolyte system by laser pulse-induced photocurrent measure­
ments [4.30]. Ambipolar diffusion coefficients (£>amb) were determined in ethanol and
actetonitrile electrolytes with varying compositions. For a given electrolyte composition,
Optimization of redox mediators and electrolytes 123

Fig. 4.3 Photocurrent transients induced by pulsed UV irradiation for 7.2-/mi thick Ti02 mes-
oporous electrodes in ethanol with 0.5 M of various salts. These included L1CIO4 (bold line),
NaC104 (solid line), TBA+C1C>4 (dotted line), and DMHI+Br (broken line). The inset shows the
amount of photogenerated electrons as a function of the pulse intensity in the presence of Li+
(circules), Na+ (squares), TBA+ (triangles), and DMHI+ (diamonds).

Damb increased with the light intensity, and the maximum of photocurrent as well as the
total photogenerated electron increased in the following order: dimethylhexylimidazo-
lium (DMHI+) < tetrabutylammonium (TBA+) < Na+< Li+, as shown in Figure 4.3. For
each cation, D^ increased by several orders of magnitude with an increasing cation
concentration. Transient photocurrent data for TBA+, and low concentrations of Li+
and Na+, accurately followed the ambipolar diffusion model. (Eq. 4.7)
At high concentrations of Li+ and DMHI+, an anomalous increase of Z>Amb was
observed. Such anomalous phenomena can be explained by the effect of the cation
adsorption and its influence on surface trap states. In the case of TBA+ and DMHI+, the
spectral analysis showed that the DMHI cations were effectively adsorbed onto the Ti0 2
film with a multilayer style, while TBA+ almost completely remained in the solution
without apparent adsorption onto the Ti0 2 film. In typical DSCs, Li+ and imidazolium
cations often coexist in the electrolytes, where Li+ adsorption on Ti0 2 film can enhance
the photo-electron generation, and the multi-adsorption of DMHI+ can increase the
electron diffusion in Ti02-electrolyte systems for high conversion efficiencies.

4.3.3 Additives
In DSC electrolytes, it has been found that the device performance can improve
with nitrogen-containing heterocyclic compounds such as ter-butylpyridine (f-BP).
124 Dye-Sensitized Solar Cells

As shown in Figure 4.4, the two photocurrent-voltage curves correspond to the two
DSCs: the solid curve represents the device in which the dyed Ti0 2 film was treated
with t-BP, while the other corresponds to the device without f-BP treatment. The treat­
ment dramatically improved the open-circuit voltage from 0.38 V to 0.66 V with­
out affecting the short-circuit current [4.2]. A similar effect was found for the other
nitrogen-containing additives, as displayed in Figure 4.5 [4.35-4.43].
The increase in Voc was due to the suppression of the back electron transfer
(Kh2) at the dyed-Ti02/electrolyte junction. Such a back transfer originates from the
reduction of triiodide by the electron on the semiconductor, and is considered to be
an essential factor for decreasing the device efficiency. The back electron transfer
occurred on the surface of Ti0 2 , unable to be covered by dye molecules, and thus trii­
odide anions were adsorbed. With addition of pyridine derivatives in the electrolyte,
the basic pyridine molecules were expected to adsorb onto the bare Ti0 2 surface due
to its Lewis acidity, thus preventing the invasion of triiodide and decreasing the rate
of undesirable electron transfer from the Ti0 2 to triiodide. A similar effect was found
for a large amount of nitrogen-containing heterocyclic compounds such as pyridine
derivatives, benzimidazole derivetives, triazole derivetives, pyrazole derivatives and
imidazole derivatives, as shown in Figure 4.5. It was also observed that the molecular
size of the derivatives was strongly related to the suppression effect. The smaller the
size of the derivatives used in the elctrolytes, the higher was the Voc of the correspond­
ing cells, which was due to the effective adsorption on Ti0 2 for smaller sizes of the
derivatives.
In addition, the increase in Voc can also be explained by the negative movement
of the Ti0 2 flatband potential (VFB) due to the adsorption of the additives. The VFB can
move negatively for the Ti0 2 electrode in the electrolyte upon the addition of the basic

Fig. 4.4 Photocurrent-voltage characteristics of a cell with an electrolyte consisting of 0.3 M


Lil + 0.03 M I2 in acetonnitrile. Dashed line: untreated dyed Ti02 electrode; solid line: same
electrode dipped for 15 min in 4-tert-butylpyridine after dye coating.
Optimization of redox mediators and electrolytes 125

Fig. 4.5 Several nitrogen-containing heterocyclic compounds used as additives in DSC


electrolytes.

derivatives. The Voc is determined by the difference of the flatband potential of Ti0 2
and the standard reduction potential Vred according to:
(4.8)
If Vred remains constant upon the addition of the derivatives, Voc will thus
increase with an increase in VFB.
The other effect on Voc may be explained due to the decrease of the triiodide
concentration in the electrolyte with the basic additives. It has been reported that
certain nitrogen-containing heterocyclic compounds such as pyrazoles can react with
triiodide in the electrolyte to form charge-transfer complexes between I and N atoms
on the pyrazole molecules [4.37].

4.3.4 Electron mediators


An electron mediator is a key material in an electrolyte, and its function is to reduce
dye cations anchored on the surface of the semiconductor, and carry electrons from
126 Dye-Sensitized Solar Cells

the counter electrode to the dyed semiconductor in DSCs, as shown in Figure 4.2.
As an electron mediator, its redox potential should be higher than that of the dye
so that the oxidized dye can be reduced in view of energy level matching. In the
DSCs, open-circuit voltage is determined by the difference in the Fermi energy of
the illuminated semiconductor and a redox couple potential. Thus, a favorable redox
potential of an electron mediator is expected to achieve a high open-circuit voltage,
while having a sufficient driving force for the reduction of oxidized dye molecules.
At present, a triiodide/iodide redox couple is the common choice in DSCs with high
conversion efficiencies. Nevertheless, this couple has its disadvantage: a significant
portion of the visible light is screened from the sensitized Ti0 2 electrode due to the
absorption of the triiodide with high concentration in the electrolyte, its low redox
potential limits the open-circuit voltage, and its corrosivity on most metals prevents
its use as a grid to collect current in large modules. Efforts to find new electron
mediators in order to replace the triiodide/iodide redox couple have been made
[4.44_4.52].
As an electron mediator in DSCs, the rate (Kn) of rereduction of the dye cation
by a redox couple should be higher than the rate of recombination of the dye cations
with electrons injected into the semiconductor. Moreover, the charge recombination
should be slow to avoid recombination between electrons on the semiconductor and
the reduced part of a redox couple in the electrolyte before electrons transfer to the
external circuit. For instance, ferrocene/ferrocenium (FeCp2+/0) has a redox potential
of 0.31 V (versus SCE), which is comparable to that of the Γ/Ι2 couple (0.15 V).
With a FeCp2/0 redox couple in an electrolyte, a certain efficiency can be expected in
view of energy level matching in the DSC. Actually, the light injected electron rate is
almost equal to that of the back electron transfer due to the latter being fast (Kh2) from
Ti0 2 to ferrocenium ions. Thus, no photocurrent can be detected in the DSC with a
FeCp^70. In the case of an electrolyte with a (SCN)2/SCN couple, the redox potential
is more positive than that of I~/T, but no improvement on the open circuit voltage in
the device has been observed. In fact, the device presented a poor efficiency in com­
parison to that of the device with the r/I'-containing electrolyte. This was caused by a
slow regeneration rate of the dye cations by the SCN~ ions. Therefore, the kinetics of
electron transfer processes related to an electron mediator may represent very impor­
tant factors for influencing the device performance, and should thus be taken into
account when developing new redox mediators. Up until now, a series of new electron
mediators have been developed for application in DSCs, but most of the devices have
presented poor conversion efficiencies, mainly due to mass-transport limitation and
unsuitable chemical kinetics related to the redox couples in the systems. Polypyridyl
cobalt (II/III) complexes and a (SeCN^/SeCN- redox couple have so far been found
to be the most promising electron mediators used in DSCs.
A polypyridyl cobalt (II/III) complex, [CoII/m(dbbip)2] (C104)2 (dbbip: 2,6-
bis(l-butylbenzimidazol-2-yl) pyridine) was recently reported to be an attractive alter­
native to the I~/l~ couple, and the photovoltaic performances are listed in Table 4.6.
In a organic solvent-based electrolyte with a mediator, a total conversion efficiency
of 7.9 % was achieved for a low light intensity ranging from 1.5-10 mW/cm2. This
result is comparable to those reported with the I~/l~ couple. At a high light intensity
of 100 mW/cm-2, however, the conversion efficiency was dramatically decreased to
Optimization of redox mediators and electrolytes 127

Table 4.6 The photovoltaic performance parameters of DSC with different redox couples in
the electrolyte at varying illumination intensities.

Redox couple Power Voc J&c FF 77 % Ref.


mW/cm 2 mV mA/cm 2

i-/i- a 9.439 722.52 1.72 0.789 10.30 1


55.331 778.56 11.45 0.771 10.73
99.762 795.65 17.77 0.748 10.58
[CoIWII(dbbip)2](C104)2 1.5 690 0.24 0.77 7.9 2
10 765 1.35 0.73 7.9
100 840 8.40 0.56 3.9
(SeCNH/SeCN-0 9.5 8.3 3
30 8.3
51.7 8.1
99.7 699 14.56 0.735 7.5
a
Nanocrystalline Ti0 2 films sensitized by black dye with an electrolyte consisting of 0.6 M dimeth-
ylpropyl imidazolium iodide, 0.1 M lithium iodide, 0.05 M iodine, and 0.5 M teri-butylpyridine in
acetonitrile [4.3].
b
In the device, nanocrystalline Ti0 2 films were sensitized by Ζ907 dye with the electrolyte containing
Co11 and Co111 complexes with concentrations of 9 x 10"2 and 1 x 10"1 M, respectively, in combination
with additives in acetonitrile/ethylene carbonate, dbbip: 2,6-bis(l-butylbenzimidazol-2-yl)pyridine;
Z907 dye: Ru(H2dcbpy)(dnbpy)-(NCS)2, where H2dcbpy is 4,4'-dicarboxylic acid-2,2'-bipyridine and
dnbpy is 4,4'-dinonyl-2,2'-bipyridine [4.51].
c
Nanocrystalline Ti0 2 films sensitized by Ζ907 dye with an electrolyte consisting of 0.15 M K(SeCN)3
0.1 M guanidinium, thiocyanate, and 0.5 M ΛΑ-methylbenzimidazole in l-ethyl-3-methylimidazolium
selenocyanate (EMISeCN) [4.47].

3.9 %. This was ascribed to slow regeneration kinetics and a limited diffusional trans­
portability of the mediator under high light illumination [4.44].
Another attractive alternative is the (SeCN^/SeCN- redox couple. With ionic
liquid electrolytes based on this (SeCN^/SeCN- couple with a low-viscosity ionic
liquid l-ethyl-3-methylimidazolium selenocyanate (EMISeCN), high conversion
efficiencies have been reported, as shown in Table 4.6. The efficiency of 7.5 % is
an inspiring value for DSCs with a nonvolatile ionic liquid electrolyte. Unlike other
electron mediators, the device with the (SeCN)3/SeCN~-based electrolyte presents a
conversion efficiency at varying light intensities ranging from 9.5 to 100 mW/cm-2,
which is similar to the device with the Γ/Ι"-containing electrolyte. This result indi­
cates that the mass transport of both (SeCN)3 and SeCN" is not limited in the device
operation. A transient absorption decay kinetic study was further carried out on the
oxidized dye on a Ti0 2 film, and the result showed that both SeCN" and Γ presented
a similar dynamic process for the dye cation regeneration. It was found that the dif­
fusion coefficients of (SeCN)3 and SeCN" in the electrolyte were 2.80 x 10~6 and
1.28 x 10~6 cm2/s, respectively. These values were higher than those for iodide and
triiodide in a pure PMII ionic liquid [4.45].
Although inspiring conversion efficiencies were reported for the DSCs with the
above-mentioned electron mediators, the long-term stability of the devices remains
unknown. Recently, it was mentioned that the long-term stability of the device with
the (SeCNg/SeGNT-based electrolyte was poor under thermal and light-soaking dual
128 Dye-Sensitized Solar Cells

stress conditions [4.53]. By now, the iodide/triiodide couple (Γ/Ι3) has been determined
to be the most efficient redox mediator, suitable for a majority of devices with different
sensitizers (metal complex dyes and organic dyes) and also with varying semiconduc­
tors, such as Ti0 2 and ZnO. In the typical I~/T system, back electron transfer from Ti0 2
of Ι~/Γ ions in the electrolyte is a slow two-electron transfer process, whereas regenera­
tion of I" from I3 is very rapid at the platinum counter electrode. Both ions have high
diffusion coefficients in the electrolyte. Table 4.6 shows the photovoltaic performance
parameters of DSCs with different redox couples I~/T, [CoII/in(dbbip)2](C104)2, and
(SeCN)3/SeCN~in the electrolyte, at varying illumination intensities.

4.4 IONIC LIQUID, QUASI-SOLID AND SOLID


ELECTROLYTES

Although stability has been manifested in low volatile organic-solvent electrolyte-


based DSCs [4.54], where interface engineering of Ti0 2 has been carried out through
co-adsorption of the amphiphilic Z-907 dye and 1-decylphosphonic acid, the seal­
ing of the device with volatile electrolytes remains a critical issue for large modules
and practical applications. An electrolyte based on room-temperature ionic liquids
has been regarded as a promising alternative, able to penetrate into the porous Ti0 2
film, and give rise to a high device stability [4.20]. Recently, ionic liquid, quasi solid
and solid state electrolytes have been developed as attractive alternatives to organic
solvent electrolytes [4.55-4.68]. This part focuses on the iodide/triiodide-based ionic
liquid, quasi-solid, and solid state electrolytes used in DSCs.

4.4.1 Ionic liquid electrolyte


Ionic liquids can afford particular advantages: a low melting point, decent chemical
and thermal stabilities, a negligible vapor pressure, non-flammability, a wide elec­
trochemical window, and a high ionic conductivity and solubility for most organic
and inorganic materials. Consequently, room-temperature ionic liquid electrolytes are
regarded as the most attractive option in DSCs for future commercial applications. As
compared to organic solvent-based electrolytes, ionic liquid electrolytes usually have
relatively high viscosities, and the charge transport of I~ is affected due to the low
diffusion transport. As a result, a higher concentration of I~ is required to compensate
for the low transport.
The performance and stability of ionic liquids for DSCs has been extensively
investigated in our group [4.55-4.59]. Figure 4.7 shows the effect of the concentration
on the viscosity and conductivity of the electrolytes, where I2 was added to various
iodide imidazohum salts [4.55]. Increases in conductivity and a decrease in viscosity
have been observed for an increasing I~ concentration in the electrolytes. The aug­
mentations of viscosity of the ILEs were found to depend on the different alkyl chain
length on the ring of imidazohum cations, which is discussed later. The diminutions of
viscosity with increasing I~ concentration in the electrolytes were explained as prob­
ably being due to the stronger delocalization of the negative I~ anions as compared
Optimization of redox mediators and electrolytes 129

Fig. 4.7 The effect of the I~ concentration in ILE on viscosity (left axis; open triangle, PMIml;
open square, HMIml; open diamond, NMIml) and conductivity (right axis; closed triangle,
PMIml; closed square, HMIml; closed diamond, NMIml), measured at 25 °C.

to that in I" in the electrolytes. In the viscous molten salts, the transfer was limited
by the diffusion of I~ ion, leading to a decrease in viscosity, giving rise to a high
conductivity.
In ILEs, the diffusion coefficient of the triiodide is about 1-2 orders of magnitude
lower than that in organic liquid electrolytes (OLEs). Thus, an I~ concentration gradient
may be produced in DSCs due to the slow diffusion of I~. This slow diffusion has been
considered a limiting factor for the device performance. Thus, a high concentration of
I3 is required to compensate for the low transport. Figure 4.8 shows the influence of
the I3 concentration on / sc as measured at AM 1.5, 100 mW cm-2 irradiance for three
iodide imidazolium-based for three iodide imidazolium-based ionic liquid electrolytes
(ILEs). All ILEs showed an increase in / sc with increasing I~ concentration at a low I~
concentration range, however / sc decreased with a further increase in I~ concentration at
high concentrations. This tendency could be explained by the slow diffusion and light
absorption of I~. The limiting current is expressed as [4.19]:
(4.9)
where P is porosity, F is Faraday's constant, DB_ is the diffusion coefficient of I~, C
is the initial concentration of I~, and w is the film thickness. The calculated values
are also shown in Figure 4.7. The values presented the same order as those measured
within the low I~ concentration range, indicating that the / S c was limited by the trans­
port of I3 to the counter electrode.
The limitation of the / S c was further supported by the experiment, in which / s c s
were measured in the cells with various thicknesses of additional electrolyte layers. The
/sc decreased dramatically for the ILE based DSCs in the thickness range 0-40 /im, while
it remained almost unchanged for the OLE-based DSCs under equivalent conditions.
130 Dye-Sensitized Solar Cells

Fig. 4.8 The effect of the I~ concentration on / sc from DSC using ILE. Measured values are
plotted as closed triangles for PMIml, closed squares for HMIml, and closed diamonds for
NMIml. Calculated / sc values from eq. 4.1 are shown with the dotted curve for PMIml, the short
dashed curve for HMIml, and the long dashed curve for NMIml. All / s c values were measured
under AM 1.5, 100 mW cm -2 irradiation.

Figure 4.9 shows / S c values measured at light intensities between 1 and


500 mW cm -2 for the DSCs with ILEs or OLEs. In OLE, the / S c increased almost
linearly with light intensity, due to the low viscosity of the organic solvent allowing
an I3 transport with high Dïy In the case of ILE (HMIml) with low I~ concentration
(<0.5 M), the / S c did not increase linearly with the light intensity as a result of a low
transportability of I~ in the viscous ILEs. / S c values under various light intensities
were also measured for PMIml and NMIml. PMIml was found to present a better
linearity for the increase of / S c as opposed to the other two ILEs. This linearity can be
explained by the difference in viscosity as discussed above.
As mentioned above, one reason for the / S c decrease was the light absorption
of I3 with the required high concentration in ILE based DSCs, as compared to OLE-
based DSCs under the same light intensity. Figure 4.10 shows absorption spectra of
ILE ([I3-] = 1 M), OLE ([I3-] = 0.05 M), and dye-adsorbed Ti0 2 . The absorbance of
these spectra was normalized to the values corresponding to 1 ßm thickness of these,
using absorption coefficients of dye or I~ (£dye = 1.26 x 104M_1 cm -1 at 530 nm, ε{3 =
2.55 x 104 M_1 cm -1 at 360 nm). As can be seen, I~ absorbed near UV to visible light
strongly, and the tail of this absorption band extended to over 500 nm.
Figure 4.11(a) shows IPCE spectra of DSCs using OLE and ILE. IPCE is
defined as the number of generated electrons divided by the number of incident pho­
tons. A strong dip in the IPCE spectra observed around 380 nm in ILE should be
attributed to the absorption of I~ in ILE, since the wavelength was identical to that of
the peak of the absorption spectrum of ILE. IPCE values for all ILEs were lower than
that of OLE over the entire wavelength range. This was not only due to the absorption
Optimization of redox mediators and electrolytes 131

Fig. 4.9 The effect of the incident light intensity on the / sc of a DSC using OLE (circles) and
ILE (HMIml, squares) with various I~ concentrations: 2 M (long dashed curve); 1 M (dashed
curve); 0.5 M (short dashed curve); 0.3 M (dotted curve).

Fig. 4.10 Absorption spectra of OLE [I~] = 0.05 M (bold curve), ILE [I~] = 1 M (bold dashed
curve), and dye adsorbed Ti0 2 film (solid curve) normalized to a l-μιη thick, conducting glass
(dotted curve); and AMI.5 photon flux density (dashed curve, right axis).

of I3 but also other factors, such as a slow I3 transfer and a high charge recombination
between the high-concentration I3 and electrons on the T i 0 2 film. An attempt was
made to extract the effect of light absorption by I3 on / s c from the following estima­
tion. The IPCE spectra for ILE were normalized to OLE by using the IPCE value at
550 nm, where the absorption of I3 was negligible. As shown in Figure 4.11(b), the
132 Dye-Sensitized Solar Cells

Fig. 4.11 (a) Observed IPCE spectra of DSC with OLE (bold solid curve) and ILE containg
IMI3 (PMIml, dotted curve; HMIml, short dashed curve; NMIml; long dashed curve), (b)
IPCE observed for OLE (bold solid curve), IPCE for ILE normalized at 550 nm (PMIml, dotted
curve; HMIml, short dashed curve; NMIml, long dashed curve), and calculated LHE for OLE
(thin solid curve) and ILE (thin dotted curve).

normalized IPCE spectra for ILE demonstrated an excellent agreement with that of
OLE over the longer wavelength range (>550 nm).
In the initial report [4.20], some ionic liquids were employed to simply replace
organic solvents in the electrolyte for DSCs. The device performance demonstrated
an outstanding stability, with an estimated sensitizer turnover exceeding 50 million.
However, the performance of the initial devices with ionic liquid electrolytes was not
satisfied, partly due to charge transport limitation in the viscous electrolyte. It is thus
essential to improve the charge transfer of I~ in ILE-based DSCs; not only by increas­
ing I3 concentration, but by decreasing the viscosity and increasing the conductivity of
the electrolytes. Usually, two ionic liquids are used in the electrolytes: one as the iodide
source, and also as a solvent, in which an imidazolium cation is usually coupled with
an electractive iodide anion, and the other as a pure solvent with low viscosity, in which
an imdazolium is coupled with other electroinactive anions. Therefore, an electrolyte
of low viscosity can be obtained by mixing both types of ionic liquids for application in
DSCs in order to improve the diffusion transportability of iodide and triiodide. Among
ionic liquid compounds, imidazolium-based ionic liquids have been extensively inves­
tigated in DSCs with high light-to-electricity conversion efficiencies.

4.4.2 Active iodide molten salts


The viscosity behavior of ionic liquids is explained as due to the interplay of coulombic
and van der Waals interaction as well as hydrogen bond formation. The iodide molten
salts present high viscosities at room temperature due to the high polarity and localized
charge density on iodide. In a special ionic liquid, the viscosity of its iodide salt can only
be changed through the modification of the cations. For l-alkyl-3-methylimidazolium
Optimization of redox mediators and electrolytes 133

iodides (alkyl chain: C3-C9), the viscosity of the molten salts was found to increase with
an increasing alkyl chain length, whereas the conductivity decreased with increasing
viscosity [4.57]. In the system, an augmentation of the alkyl chain length can decrease
the electrostatic attraction between the imidazolium cations and iodide anions in one
way, but the van der Waals interaction increases, and finally gives rise to an increase in
viscosity with the augmentation of the alkyl chain length. l-Ethyl-3-methylimidazolium
iodide (DMII, Mp: 77.5 °C) and 1,3-dimethylimidazolium (DMII, Mp: 92 °C), how­
ever, are solid at ambient temperature due to their high lattice Gibbs energies in their
rigid conformation with small and symmetric cations. Among all the room-temperature
ionic liquid iodide imidazolium salts, l-propyl-3-methylimidazolium iodide (PMII)
presents the lowest viscosity (1084 cP) and highest conductivity (0.54 mS cm-1) at
room temperature, and has for this reason been widely used in DSCs with ionic liquid
electrolytes. According to reports, an efficient DSC is one with an ionic liquid-based
electrolyte in combination with an amphilic polypyridil ruthenium sensitizer, Z907,
as shown in Figure 4.13, in which the ionic electrolyte comprised 0.5 M iodine and
0.45 M N-methyl-benzimidazole (NMBI) in pure PMII [4.62]. An efficiency of 6.6 %
was achieved for DSCs with single pure iodide molten salt as both the iodide source
and solvent at 100 mW/cm2. Up until now, imidazolium iodides have been extensively
employed for DSCs in both OLE and ILE, and their chemical structures are shown in
Figure 4.12.
It was recently reported that viscosity and conductivity of ionic liquid-based
electrolytes can be adjusted through the simple mixing of two or three molten iodide
salts. For instance, EMU and DMII are solid salts with melting points of 77.5 and
92 °C, respectively. The eutectic salt shows a higher conductivity and a lower melt­
ing point of 47.5 °C at a ratio of Γ/1~ as compared to both EMU and DMII [4.53].
The low melting point can be explained by the increase in entropy of the components
in the mixture. By mixing EMU, DMII, and l-allyl-3-methyl imidazolium iodide

Fig. 4.12 Chemical structures of various iodide ionic liquids used in DSCs.
134 Dye-Sensitized Solar Cells

Fig. 4.13 Chemical structures of amphiphilic ruthenium complexes: N3, Z907, C103, and
K19dyes.

(Mp: 60 °C), the prepared ternary melt has a melting point below 0 °C with a strikingly
high conductivity (σ) of 1.68 mS. This is higher than that of PMII (0.58 mS cnr 1 ).
The ternary salt was used as the ionic source and solvent in an electrolyte, which
also contained other components of iodine and two additives: Λ^-butylbenzoimidazole
(NBB) and guanidinium thiocyanate (GNCS). With this electrolyte, a power conver­
sion efficiency of 7.0 % was achieved in the DSCs, in which Ti0 2 was sensitized with
Z-907Na in combination with 3-phenylpropicnic acid (PPA) [4.53].
The use of an electrolyte with a high concentration of imidazolium iodide in a
DSC gives rise to an undesirable quenching of the dye molecules according to [4.64]:

In DSCs with an electrolyte containing high concentrations of iodide, the


excited dye molecules are unable to inject electrons into the Ti0 2 conduction band.
They can however accept electrons from Γ to form a reduced D", which no longer
Optimization of redox mediators and electrolytes 135

injects electrons into the semiconductor. It has been reported that 25 % of Z907Na
dye molecules undergo the reductive quenching reaction in the electrolyte in the
presence of pure PMII. Therefore, in ILE-based DSCs, it is essential to reduce the
iodide concentration by mixing an active iodide ionic liquid with a nonactive ionic
liquid to prevent undesirable quenching of the dye molecules. Moreover, a low vis­
cosity and a high conductivity may be realized by addition of a low viscous nonactive
ionic liquid.

4.4.3 Nonactive iodide molten salts


Recently, some electroinactive imidazolium salts with anions such as thiocyanate and
dicyanoamide have demonstrated very low viscosities at ambient temperature. With
these ionic liquids in the electrolytes, the power conversion efficiencies have been
dramatically improved, as compared to the devices with pure iodide ionic liquid as
solvent. In recent years, great effort has been directed toward the development and
investigation of various low-viscosity electroinactive ionic liquids used as solvents
for the application in ionic liquid-based electrolytes. The low-viscosity of non-active
ionic liquid is essential for the improvement of the charge diffusion transport in the
electrolyte, and the dilution of iodide salt by the viscous ionic liquid can effectively
prevent the undesirable quenching reaction due to the high concentration of iodide
in the pure active iodide salt-based electrolyte as mentioned above. A series of elec­
troinactive imidazolium ionic liquids with various non-aromatic anions have been
developed and investigated. These anions include thiocyanate, dicyanoamide, tricyy-
anoamide, tetracyanoborate, etc., as shown in Figure 4.14. As compared to the pure
iodide melts, these ionic liquids demonstrated much lower viscosities at the ambient
temperature, which can be ascribed to the reduced ion-paring effect due to the strong
charge delocalization on the delocalized anions.
Two efficient DSCs have been reported on: one with a low-volatile electrolyte
and the other with an ionic liquid electrolyte, both in combination with C103 dye (as
show in Fig. 4.13) [4.4]. The composition of the low-volatile electrolyte in device
A was the following: 1.0 M 1,3-dimethylimidazolium iodide (DMII), 0.15 M I2,
0.5 M Λ^-butylbenzoimidazole (NBB), and 0.1 M guanidinium thiocyanate (GNCS)
in 3-methoxypropionitrile (MPN). For the ionic liquid electrolyte in device B it was:
DMII/1 -ethy 1-3-methylimidazolium iodide/1 -ethyl-3-methylimidazolium tetracy-
anoborate/I2/NBB/GNCS (molar ratio: 12/12/16/1.67/3.33/0.67). Under AM 1.5G
conditions (99.8 mW cm-2), the short-circuit photocurrent density (/sc), open-circuit
photovoltage (V0CX and fill factor (FF) of device A with a low-volatility, 3-methoxy-
propionitrile-based electrolyte were 17.51 mA cm -2 ,771 mV, and 0.709, respectively,
yielding an overall conversion efficiency (77) of 9.6 %. Under equivalent conditions,
the photovoltaic parameters (/sc, Voc, FF, and 77) of device B with the ionic liquid
electrolyte were 15.93 mA cm-2, 710 mV, 0.747, and 8.5 %, respectively. Under a
low light intensity of 9.4 mW cm-2, a very impressive efficiency of 9.1 % could be
achieved, as displayed in Figure 4.15. The highest values were obtained in DSCs with
sol vent-free ionic liquid electrolytes.
The long-term stability test was carried out for the two devices as presented in
Figure 4.16. Both devices with a low-volatility electrolyte and a solvent-free ionic
136 Dye-Sensitized Solar Cells

Fig. 4.14 Chemical structures of several non-active ionic liquids used in DSCs.

Fig. 4.15 I-V characteristics of device B with the C103 sensitizer and a solvent-free ionic liq­
uid electrolyte measured (a) in the dark and (b-d) under an irradiance of AM 1.5G sunlight
(100 mW cm-2), (b) 9.4 mW cm"2; (c) 51.3 mW cm"2; (d) 99.8 mW cm"2. The aperture area of
the metal mask was 0.158 cm2. A UV-absorbing antireflection film covered the cell during the
measurements.
Optimization of redox mediators and electrolytes 137

Fig. 4.16 Detailed photovoltaic parameters measured under an irradiance of AM 1.5G sunlight
for devices A and B during successive full sunlight soaking at 60 °C.
138 Dye-Sensitized Solar Cells

liquid electrolyte showed good stabilities, maintaining 91 % and 94 % of their initial


efficiencies of 9.6 % and 8.4 %, respectively.
Apart from the electroinactive imidazolium iodide salts, other cations such as
ammonium, phosphonium, sulphonium, pyridinium, pyrrolidinium and guanidinium,
can form room-temperature ionic liquids with iodide. These ionic liquids have also
been investigated in DSCs, and their structures are given in Figure 4.17.
Binary melts of s-ethyltetrahydrothiophenium iodide and dicyanoamide (or tri-
cyanomethide) have been synthesized and employed in DSCs, and high power conver­
sion efficiency up to 6.9 % have been achieved in combination with the Z907Na dye
under the illumination of AM1.5G full sunlight [4.69]. This value is the highest meas­
ured if one excludes DSCs with imidazolium-based ionic liquids. The transport of
triiodide was determined at a high iodide concentration, and the viscosity-dependence
was explained by a physical diffusion-coupled bond exchange mechanism differing
from the simple physical diffusion. It was found that certain counter anions in the
electrolytes, e.g., dicyanoamide, have a significant influence on surface states and
electron transport in Ti0 2 films.
A series of tetralkylamonium iodides, NR2R2 (R = methyl or ethyl, and
R = pentyl or hexyl), have been synthesized and incorporated as both the iodide source
and the solvent in DSCs. Under illumination with a low light intensity of 10 mW cm-2,

Fig. 4.17 Chemical structures of cations of ionic liquids used in DSCs.


Optimization of redox mediators and electrolytes 139

an optimized conversion efficiency of 2.4 % was obtained for the electrolyte with
the room-temperature ionic liquid of (Me2Hex2N)I in combination with TBP as the
additive. For the soft solid electrolyte with (Et2Hex2N)I, the corresponding value was
2.3 % under identical conditions [4.70]. Asymmetric aliphatic tetraalkyl phosphonium
iodides of varying alkyl chain lengths were synthesized and investigated as the iodide
source in electrolytes with organic solvents. The conversion efficiency was found to
be 5.9 % at a 8.9-mW cnr 2 light intensity [4.71].
Quarternary ammonium iodides with various alkyl chain lengths have been
employed as the iodide source in electrolytes. With tetrahexylammonium iodide, a
retardation of the interfacial charge transfer at Ti02/dye/electrolyte was observed in
comparison to Li+ and other ammonium iodides with short alkyl chain lengths. It was
found that a further increase in the alkyl chain length of the ammonium salt did not
improve the electron lifetime on Ti0 2 , but rather decreased the fill factor. This pointed
at the fact that THA+ was optimized in size with regard to the ammonium cation,
giving rise to highly efficient DSCs [4.72].

4.4.4 Additives in ILEs


Several additives have been employed in ionic liquid-based electrolyte to enhance
the photovoltaic performance. In order to improve the photovoltage of DSCs, some
nitrogen-containing heterocycles have been utilized as Lewis bases in electrolytes.
Routine additives include 4-ter-butylpyridine and Af-alkylbenzoimidazole. These
basic additives have been reported to adsorb onto the Ti0 2 surface, and also to shift
the conduction band negatively, resulting in an improvement of the photovoltage. In
addition, the adsorption of these additives onto the Ti0 2 surface is believed to passi-
vate recombination centers and thus reduce back electron transfer from Ti0 2 to I~.

4.4.5 Quasi-solid electrolyte


An ionic liquid-based electrolyte is one of the most attractive charge mediators used
in DSCs for practical applications. Solidification of the liquid state electrolytes may
be necessary in DSCs, especially when considering applications for matching differ­
ent architectural constructions. Quasi- or solid-state electrolytes can be realized by
mixing the liquid electrolytes with polymers, inorganic silica nanoparticles, and low
molecular weight organic gelators. There exist some reports on the solidification of
volatile organic solvent-based electrolytes in DSCs in order to prevent the solvent
leakage and evaporation, expected to improve the device stability. It should be noted
that the solidification only affects their macroscopic fluidity, and can not radically
solve the issue of evaporation and leakage in the DSCs, especially under thermal
stress. This is due to the solidification not being able to significantly improve boiling
points of the solvents used in the electrolytes. The present section thus focuses on
sol vent-free electrolytes.
Low molecular weight organic gelator have been extensively employed to solid­
ify ionic liquid-based electrolytes due to their physical thermoreversibility. Figure 4.18
shows the chemical structures of some common gelators reported in DSCs. With these
gelled electrolytes, dyed nanoporous Ti0 2 films can be effectively filled above the
140 Dye-Sensitized Solar Cells

Fig. 4.18 Chemical structures of small molecular gels used in DSCs.

sol-gel transition temperature (Tgel), and a mechanically stable quasi-solid electrolyte is


obtained upon cooling. The characteristics of the gelators originate from their superior
ability to form intermolecular hydrogen bonds between the oxygen atom in the ure-
thane group and the hydrogen atom in the amide group. In addition, long alkyl chains
in the molecules also contribute to form stiff self aggregates via van der Waals forces.
Our previous report [4.56] describes the preparation of quasi-solid electrolytes
by mixing 40 g L_1 of gelator 1 (shown in Fig. 4.18) with various l-alkyl-3-methyl
imidazolium iodides (alkyl: C3-C9). The overall conversion efficiency of 5.0 % was
obtained from a gel electrolyte with l-hexyl-3-methylimidazolium iodide (HMII).
The photovoltaic parameters were comparable to those for the DSCs without the gela­
tor in the ionic liquid electrolyte under comparable conditions [4.73].
A quasi-solid state electrolyte has also been prepared by iodide ionic liquid
crystal without any gelling materials [4.58, 4.59]. It was reported that a novel ionic
liquid crystal (ILC) system (Ci2MImI/I2) with a smectic A phase used as an elec­
trolyte for a DSC showed the highest short-circuit current density and the best light
to-electricity conversion efficiency when compared to a system using the non-liquid
crystalline ionic liquid (CnMImI/I2). This was due to an assembled structure of the
imidazolium cations in the crystal with a high conductivity. In this system, the larger
exchange reaction-based diffusion coefficients (Dex) were observed to support the
higher conductivity of ILC due to an enhancement of the exchange reaction between
iodide species.
Another similar structural gelator (denoted 2 in Fig. 4.18) was employed to
obtain a mixture with a low-viscosity binary ionic liquid blend of l-propyl-3-methyl-
imidazolium iodide (PMII) and l-ethyl-3-methylimidazolium thiocyanate (EMINCS)
Optimization of redox mediators and electrolytes 141

for the preparation of a quasi-solid electrolyte. With 2 wt % of the gelator in the


electrolyte, the r gel was 108 °C, and well-organized supermolecular structures were
observed. The electrolyte in combination with an amphiphilic ruthenium sensitizer
K-19 in DSCs, gave rise to a short-circuit photocurrent, open-circuit voltage and fill
factor of the device of 12.8 mA/cm2, 706 mV and 0.72, respectively, yielding an over­
all power conversion efficiency of 6.3 %. These values were similar to those obtained
for DSCs without gelator in the electrolyte [4.69].
Solidification of the electrolyte can also be realized by dispersing the chemi­
cally and electrochemically stable inorganic nanoparticles. A quasi-solid electrolyte
was prepared by mixing 5 wt % of fumed silica nanoparticles in an ionic liquid elec­
trolyte consisting of 0.5 M iodine and 0.45 M NMBI in pure MPII. With the electro­
lyte, the short-circuit photocurrent, open-circuit voltage and fill factor of the device
were 12.75 mA/cm2, 672 mV and 0.709, respectively, yielding an overall power con­
version efficiency of 6.1 % at AM 1.5 sunlight. Under equivalent conditions, almost
identical results were obtained for the devices with the corresponding ionic liquid
electrolyte, indicating that the presence of silica nanoparticles had no adverse effects
on the conversion efficiency [4.64].
A titaniafillerwas used to solidify an electrolyte containing poly(ethylene oxide)
and an I~/T redox couple. The introduction of the titania filler into the poly(ethylene
oxide) was able to reduce the crystallinity of the polymer and enhance the mobil­
ity of the I~/T redox couple, resulting in an overall conversion efficiency of 4.2 %
under illumination at a 65.6-mW cm -2 light intensity [4.74]. A similar technique was
employed to solidify a polymer electrolyte with fumed silica nanoparticles. A three-
dimensional network was reported to be created with the introduction of the amor­
phous silica nanoparticles. It presented a surface area of 255 m2 g_1, and an efficiency
of 4.5 % was achieved at 100-mW cm -2 light intensity [4.75].
In DSCs with an ionic liquid electrolyte, a back electron reaction from Sn0 2
to I3 should be considered due to the high concentration of I~ anion in the electrolyte.
A blocking layer deposited between the conducting layer of a glass substrate and a Ti0 2
layer may be regarded as an effective way of retarding the reaction and of improving
the open-circuit voltage and overall conversion efficiency of the device [4.76, 4.77]. It
has been reported that Nb 2 0 5 films may be used as attractive blocking layers in DSCs
with ionic liquid electrolytes. The efficiency has been seen to increase from 2.8 % to
4.8 % after the introduction of a Nb 2 0 5 layer in DSCs with an ionic liquid electrolyte.

4.5 REMARKS AND PROSPECTS

As a key part of a dye-sensitized solar cell (DSC), the electrolyte strongly relates to
the photovoltaic performance and stability of the device. An electron mediator in an
electrolyte is typically used to reduce dye cations, and carry electrons from the coun­
ter electrode to the dyed semiconductor in DSCs, conducting a series of physical and
chemical processes. There is no doubt that an electron mediator is of crucial impor­
tance in an electrolyte for stable operation and high performance of a dye-sensitized
solar cell. Up until now, the iodide/triiodide couple (Γ/Ι3) has been determined as
142 Dye-Sensitized Solar Cells

the most efficient redox mediator, suitable for most devices with varying sensitizers
(metal complex dyes and organic dyes) and also with different semiconductors, such
as Ti0 2 and ZnO. Although inspiring conversion efficiencies have been reported in
DSCs with other electron mediators, these conversion efficiencies and device stabili­
ties remain unsatisfactory. In spite of this, new electrolytes are intensively expected
to further improve the light-to-electricity conversion efficiency in consideration of the
potential space in DSCs.
In combination with a volatile electrolyte, the DSC has reached respectable high
efficiencies of 11.0-11.3 % measured under the standard air mass 1.5 global (AM 1.5 G)
sunlight. In view of the practical application, solvent-free ionic liquid electrolytes may
represent an attractive option due to their particular advantages. Furthermore, solidi­
fication of the liquid state electrolytes may be necessary in DSCs, considering their
application to match various architectural constructions.

4.6 REFERENCES

[4.1] O'Regan, B., Grätzel, M., Nature 1997, 353, 737.


[4.2] Nazeerudin, M. K., Kay, A., Rodico, I., Humphry-Baker, R., Müller, E., Lisk, R, Valchopoulos, N.,
Grätzel, M., /. Am. Chem. Soc. 1993,775, 6382.
[4.3] Chiba, Y., Islan, A., Watanabe, Y., Komiya, R., Koida, N., Han, L., Jpn. J. App. Phys. 2006, 45,
L638.
[4.4] Shi, D., Pootrakuchote, N., Li, R., Guo, J., Wang, Y, Zakeeruddin, S. M., Grätzel, M., Wang, P.,
/. Phy Chem. C. 2008, 772, 17046.
[4.5] Murakoshi, K., Kogure, R., Wada, Y, Yanagida, S., Chem. Lett. 1997, 471.
[4.6] Xia, J. B., Masaki, N., Lira-Cantu, M., Kim, Y, Jiang, K. J., Yanagida, S., /. Am. Chem. Soc, 2008,
130, 1258.
[4.7] Bach, U., Lupe, D., Comet, P., Moser, J. E., Weissortel, F., Salbeck, J., Spreitzer, H. Grätzel, M.,
Nature, 1998, 395, 583.
[4.8] Tennakone, K., Kumara, G. R., Kottegoda, I. R., Pereva, V. S. P., Chem. Commun. 1999, 15.
[4.9] Sayama, K., Suguhara, H., Arakawa, H., Chem. Mater. 1998,10, 3825.
[4.10] Jiang, K. J., Masaki, N., Xia, J. B. Noda, S., Yanagida, S., Chem. Commun. 2006, 2460.
[4.11] Tachibana, Y, Moser, J. E., Grätzel, M., Klug, D. R., Durrant, J. R., /. Phys. Chem. B 1996, 100,
20056.
[4.12] Asbury, J. B., Ellingson, R. J., Ghosh, H. N., Ferrere, S., Nozik, A. J. Lian, T., /. Phys. Chem. B
1999,100,3110.
[4.13] Benkö, G., Kallioinen, J., Korppi-Tommola, J. E. I., Yartsev, A. P., Sundström, V. J., /. Am. Chem.
Soc. 2002, 724, 489.
[4.14] Wenger, B., Grätzel, M., Moser, J. E. /. Am. Chem. Soc. 2005, 727, 12150.
[4.15] Tachiba, Y, Haque, S. A., Mercer, I. P., Moser, J. E., Klug, D. R., Durrant, J. R. /. Phys. Chem. B,
2001,105, 7424.
[4.16] Schlichthrl, G., Huang, S. Y, Sprague, J., Frank, A. J., /. Phys. Chem. B 1997,101, 8141.
[4.17] Zaban, A., Ferrer, S., Gregg, B. A., /. Phys. Chem. B 1997, 702, 452.
[4.18] Fisher, A. C., Peter, L. M., Ponomarev, E. A., Walker, E. A., Wijayantha, K. G. U., /. Phys. Chem.
B, 2000,104, 949.
[4.19] Papageorgiou, N., A., Gratzel, Infelta, P. P., Sol. Energy Mater. Sol. Cells. 1996, 44, 405.
[4.20] Papageorgiou, N., Athanassov, Y, Armand, M., Bonhote, P., Pettersson, H., Azam, A., Grätzel, M.,
/. Electrochem. Soc. 1996,143, 3099.
[4.21] Kawano, R., Watanaba, M., Chem. Commun. 2003, 330.
[4.22] Enright, B., Redmond, G., Fitzmaurice, D., /. Phys. Chem. 1994, 98, 6195.
[4.23] Kebede, Z., Lindquist, S. E., Sol. Energy Mater. Sol. Cells 1999, 57, 259.
Optimization of redox mediators and electrolytes 143

[4.24] Hara, K., Horiguchi, T., Kinoshita, T., Sayama, K., Arakawa, H., Sol. Energy Mater. Sol. Cells 2001,
70, 151.
[4.25] Fukui, A., Komiya, R., Yamanaka, R., Islam, A., Han, L. Y., Sol. Energy Mater. Sol. Cells 2006,
90, 649.
[4.26] Lagemaat, J. V. D., Frank, A. J., /. Phys. Chem. B 2001,105, 11194.
[4.27] Kopidakis, N., Schiff, E. A., Park, N. G., Lagemaat, J. V. D., /. Phys. Chem. B 2000,104, 3930.
[4.28] Redmond, G., Fitzmaurice, D. /. Phys. Chem. 1993, 97, 1426.
[4.29] Liu, Y, Hagfeldt, A., Xiao, X. R., Lindquist, S. E., Sol. Energy Mater. Sol. Cells 1998, 55, 267.
[4.30] Kambe, S., Nakade, S., Kitamura, T., Wada, Y, Yanagida, S., /. Phys. Chem. B 2002,106, 2967.
[4.31] Nakade, S., Kambe, S., Kitamura, T., Wada, Y, Yanagida, S., /. Phys. Chem. B. 2001,105, 9150.
[4.32] Nakade, S., Kanzaki, T., Kubo, W., Kitamura, T., Wada, Y, Yanagida, S., /. Phys. Chem. B. 2005,
109, 3480.
[4.33] Nakade, S., Makimoto, Y, Kubo, W., Kitamura, T., Wada, Y, Yanagida, S., /. Phys. Chem. B. 2005,
109, 3488.
[4.34] Watson, D. F , Meyer, G. J., Coordination Reviews 2004, 248, 1391.
[4.35] Kusama, H., Arakawa, H., /. Photchem. Photobio. A: Chem. 2004,164, 103.
[4.36] Kusama, H., Arakawa, H., /. Photchem. Photobio. A: Chem. 2004,165, 157.
[4.37] Kusama, H., Kurashige, M., Arakawa, H., /. Photchem. Photobio. A: Chem. 2005,169, 169.
[4.38] Kusama, H., Arakawa, H., Sugihara, H., /. Photchem. Photobio. A: Chem. 2005, 777, 197.
[4.39] Kusama, H., Arakawa, H., /. Photchem. Photobio. A: Chem. 2004,162, 441.
[4.40] Kusama, H., Koshinari, Y, Sugihara, H., Arakawa, H., Sol. Energy Mater. Sol. Cells 2003, 80, 167.
[4.41] Kusama, H., Koshinari, Y, Sugihara, H., Arakawa, H., Sol. Energy Mater. Sol. Cells 2005, 85, 333.
[4.42] Hara, K., Nishikawa, T., Kurashige, M., Kawauchi, H., Kashima, T., Sayama, K., Aika, K., Arakawa, H.,
Sol. Energy Mater. Sol. Cells 2005, 85, 21.
[4.43] Fischer, A., Petttersson, H., Hagfeldt, A., Boschloo, G., Kloo, L., Gorlov, M., Sol. Energy Mater.
Sol. Cells 2007, 91, 1062.
[4.44] Nusbaumer, H., Zakeemddin, S. M., Moser, J.-E., Grätzel, M., Chem. Eur. J. 2003, 9, 3756.
[4.45] Wang, P., Zakeemddin, S. M., Moser, J.-E., Humphy-Baker, R., Grätzel, M., /. Am. Chem. Soc.
2004,126, 7164.
[4.46] Nazeeruddin, M. K., Pechy, P., Renouard, T., Zakeemddin, S. M., Humphry-Baker, R., Comte, P.,
Liska, P., Cevey, L., Costa, E., Shklover, V., Spiccia, L., Deacon, G. B., Bignozzi, C. A., Grätzel,
M., /. Am. Chem. Soc. 2001,123, 1613.
[4.47] Wang, P., Zakeemddin, S. M., Moser, J. E., Humphry-Baker, R., Grätzel, M., /. Am. Chem. Soc.
2004,126, 7164.
[4.48] Nusbaumer, H., Moser, J. E., Zakeemddin, S. M., Nazeeruddin, M. K., Grätzel, M., /. Phys. Chem.
B. 2001, 705,10461.
[4.49] Oskam, G., Bergeron, B. V., Meyer, G. J., Searson, P. C , /. Phys. Chem. B. 2001,105, 6867.
[4.50] Sapp, S. A., Elliott, C. M., Contado, C , Caramon, S., Bignozzi, C. A., /. Am. Chem. Soc. 2002,
724,11215.
[4.51] Nusbaumer, H., Zakeemddin, S. M., Moser, J. E., Gratzel, M., Chem. Eur. J. 2003, 9, 3756.
[4.52] Cameron, P. J., Peter, L. M., Zakeemddin, Grätzel, M., Coordination Chemistry Reviews 2004,
248, 1447.
[4.53] Cao, Y, Zhang, J., Bai, Y, Li, R., Zakeemddin, S. M., Grätzel, M., Wang, P., /. Chem. Phys. B.
2008, 772, 13775.
[4.54] Wang, P., Zakeemddin, S. M., Moser, J. E., Nazeeruddin, M. K., Sekigucgi, T., Grätzel, M., Nature.
Mater. 2003, 2, 402.
[4.55] Kubo, Kambe, S., W, Nakade, S., Kitamura, T., Hanabusa, K., Wada, Y, Yanagida, S., /. Chem.
Phys. B. 2003,107, 4374.
[4.56] Kubo, Murakoshi, K., W., Kitamura, T., Yoshida, S., Hamki, M., Hanabusa, K., Shirai, H., Wada,
Y, Yanagida, S., /. Chem. Phys. B. 2001,105, 12809.
[4.57] Kubo, W., Kitamura, T., Hanabusa, K., Wada, Y, Yanagida, S., Chem. Commun. 2002, 374.
[4.58] Yamanaka, N., Kawano, R., Kubo, W., Kitamura, T., Wada, Y, Watanabe, M., Yanagida, S., Chem.
Commun., 2005, 740.
[4.59] Yamanaka, N., Kawano, R., Kubo, W, Masaki, N., Kitamura, T., Wada, Y, Watanabe, M., Yanagida, S.,
/. Phys. Chem. B, 2007, 777, 4763.
144 Dye-Sensitized Solar Cells

[4.60] Matsumoto, H., Matsuda, T., Tsuda, T., Hagiwara, R., Ito, Y, Miyazaki, Y, Chem. Lett. 2001, 26.
[4.61] Wang, P., Zakeemddin, S. M., Exnar, I., Grätzel, M., Chem. Commun. 2002, 2972.
[4.62] Wang, P., Zakeemddin, S. M., Moser, J.-E., Grätzel, M., /. Phys. Chem. B 2003,107, 13280.
[4.63] Wang, P., Zakeemddin, S. M., Humphry-Baker, R., Grätzel, M., Chem. Mater. 2004,16, 2694.
[4.64] Wang, P., Wenger, B., Humphry-Baker, R., Moser, J.-E., Teuscher, J., Kantlehner, W., Mezger, J.,
Stoyanov, E. V., Zakeemddin, S. M., Grätzel, M., /. Am. Chem. Soc. 2005, 727, 6850.
[4.65] Kuang, D., Wang, P., Ito, S., Zakeemddin, S. M., Grätzel, M., /. Am. Chem. Soc, 2006,128, 7732.
[4.66] Kato, T., Okazaki, A., Hayase, S., Chem. Commun. 2005, 363.
[4.67] Zistler, M., Wächter, P., Schreiner, C , Fleischmann, M., Gerhard, D., Wasserscheid, P., Hinsch, A.,
Goresa, H. J., /. Electrochem. Soc, 2007,154, B925.
[4.68] Gorlov, M., Pettersson, H., Hagfeldt, A., Kloo, L., Inorg. Chem. 2007, 46, 3566.
[4.69] Xi, C , Cao, Y, Cheng, Y, Wang, M., Jing, X., Zakeemddin, S., Grätzel, M., Wang, P., /. Phys.
Chem. C„2008, 772, 11063.
[4.70] Santa-Nokki, H., Busi, S., Kallioinen, J., Lahtinen, M., Korppi-Tommola, J., /. Photochem.
Photobiol, A, 2007,186, 29.
[4.71] Cai, N., Zhang, J., Zhou, D., Yi, Z., Guo, J., Wang, P., /. Phys. Chem. C 2009,113, 4215.
[4.72] Kanzaki, T., Nakade, S., Wada, Y, Yanagida, S., /. Photochem. Photobiol. Sel, 2006, 4, 389.
[4.73] Mohmeyer, N., Kuang, D., Wang, P., Schmidt, H., Zakeemddin, S. M., Grätzel, M., /. Mater. Chem.
2006,16, 2978.
[4.74] Usui, H., Matsui, H., Tanabe, N., Yanagida, S., /. Photochem.Photobiol. A, 2004,164, 97.
[4.75] Kim, J., Kang, M.-S., Kim, Y. J., Won, J., Park, N.-G., Kang, Y. S., Chem. Commun. 2004, 1662.
[4.76] Xia, J., Masaki, N., Jiang, K., Wada, Y, Yanagida, S., Chem. Lett. 2006, 35, 252.
[4.77] Xia, J., Masaki, N., Jiang, K., Yanagida, S., /. Photochem. Photobiol. A 2007,188, 120.
CHAPTER 5

PHOTOSENSITIZATION OF Sn0 2
AND OTHER OXIDES
Prashant V Kamat

Nanostructure assemblies provide novel opportunities in the design of next genera­


tion's solar cell [5.1, 5.2]. The high porosity of mesoscopic oxide films which ena­
bles the incorporation of sensitizing dyes in large concentrations has facilitated the
development of dye-sensitized solar cells (DSCs) and the construction of solar panels
[5.3-5.6]. Although most of the dye sensitization work centers around nanostructured
Ti0 2 films modified with a ruthenium polypyridyl complex, numerous studies have
focused on utilizing semiconductor oxide substrates other than Ti0 2 . For example,
because of the low-lying conduction band (£ C B = 0 V VS. NHE), Sn0 2 can capture
electrons from many of the sensitizers that are generally unable to sensitize Ti0 2
semiconductors (£ C B = -0.5 V vs. NHE). Sn0 2 and ZnO films have been used by
numerous researchers to demonstrate several salient features of the dye sensitization
process. A few aspects of dye sensitization of Sn0 2 and other oxide semiconductors
are discussed in this chapter.
As shown in the previous chapters, when a dye-loaded semiconductor film is
illuminated with visible light, the sensitizer molecules absorb light and inject elec­
trons into the semiconductor particles (Fig. 5.1). These electrons are then collected
at the conducting glass surface to generate an anodic photocurrent. The redox cou­
ple (e.g., Ι3/Γ) present in the electrolyte quickly regenerates the sensitizer [5.7].
By choosing an appropriate sensitizer, it is thus possible to tune the photoresponse
of these nanostructured semiconductor films. For instance, sensitizing dyes such as
chlorophyll analogues [5.8-5.11], squaraines [5.12, 5.13], rhodamine [5.14], peryl-
enes [5.15], and oxazines [5.16], can extend the photoresponse of Sn0 2 films to the
red-infrared region. Recently, excited-state interaction between an Ru(II) polypy­
ridyl complex and CdSe has been investigated for the hole transfer process [5.17].
146 Dye-Sensitized Solar Cells

Fig. 5.1 The principle of a dye-sensitized solar cell. The scheme shows charge injection from
an excited sensitizer (S*) into semiconductor electrolytes.

5.1 DEPENDENCE OF THE SENSITIZATION EFFICIENCY


ON THE ENERGY DIFFERENCE

The energy difference between the conduction band of the semiconductor and the
oxidation potential of the excited sensitizer is the major driving force for the excited
state charge transfer [5.18, 5.19]. Since the early 1980's, various approaches have
been considered for investigating the energy gap dependence of the photosensitization
efficiency. Hashimoto et al. [5.20, 5.21] have shown that the excited-state lifetime of
Ru(bpy)3+ adsorbed onto a metal oxide semiconductor is dependent on the conduction
band energy of the semiconductor. An effort was also made to establish the energy
gap dependence of the electron transfer between silver halides and J-aggregates of the
dye [5.22, 5.23]. Spitler and his coworkers [5.24, 5.25] have varied the pH to explore
the energetic threshold for dye-sensitized photocurrent generation at SrTi0 3 and Ti0 2
electrodes. The energy gap is dependent on pH since the conduction band of metal
oxide semiconductor shifts 0.059 V/pH unit.
The charge injection from an excited sensitizer into semiconductor nanoparticles
is an ultrafast process occurring on the timescale of femto- to nanoseconds [5.26-5.40].
Electron transfer kinetics in dye-sensitized Sn0 2 and Ti0 2 systems has been evaluated
in terms of the Marcus theory [5.41, 5.42], and Hupp and coworkers [5.43] have suc­
cessfully applied this theory to probe the recombination of conduction band electrons
from Sn0 2 to an oxidized sensitizer. Other research groups [5.44, 5.45] have also
employed this model to investigate the charge recombination kinetics. As the driving
force, AG, increases, so does the rate of ET, reaching a maximum when the driving
force equals the reorganization energy.
It is known that the photosensitization efficiency of a dye-modified semiconductor
electrode is strongly dependent on the applied bias [5.19, 5.46]. It significantly decreases
when the applied potential is close to the flat band potential of the semiconductor. Often,
such a decrease is attributed to the increase in reverse electron transfer arising from the
Photosensitization of Sn02 and other oxides 147

lack of potential gradient within the semiconductor to drive away the injected charge
towards the collecting surface.
Spectroelectrochemical measurements of metal oxide films have shown that an
externally applied electrochemical bias causes electron accumulation in semiconductor
nanocrystallites [5.47-5.53]. The onset potential at which the electron accumulation is
seen corresponds to the flat band potential of the semiconductor.
In the case of an InP semiconductor, the applied potential has been shown to
influence the hot electron injection process. Several researchers have observed an
increase in the quenching of the excited state of the sensitizer adsorbed onto an n-type
semiconductor electrode or increased production of oxidized sensitizer by biasing the
electrode at positive potentials [5.24, 5.54, 5.55]. Resonance Raman spectroscopy
[5.56] and transient absorption spectroscopy [5.57] have been employed to moni­
tor the changes that occur on the nanocrystalline semiconductor surface at positive
and negative bias potentials. O'Regan et al. [5.58] have investigated the influence of
externally applied bias on the charge injection efficiency and reverse electron transfer
process in a Ti02/Ru(II) system.

5.2 COUPLED SEMICONDUCTOR SYSTEMS

In order to improve charge separation, it is possible to employ composite semiconduc­


tor films consisting of two or more semiconductors with favorable energetics [5.11,
5.59]. Up to a tenfold enhancement in the photocatalytic degradation rates has been
obtained using Sn0 2 /Ti0 2 composite films, as demonstrated in a recent study [5.60,
5.61]. The coupling of a large bandgap semiconductor with a smaller one not only
extends the photoresponse into the visible region but also facilitates charge separation
by accumulating electrons and holes in separate particles. An improved charge sepa­
ration in Ti02/CdS [5.62, 5.63] Sn02/CdS [5.64] ZnO/ZnS [5.65, 5.66] and ZnO/CdS
[5.59] has been demonstrated with transient absorption measurements. Similarly, a
charge rectification effect has been observed when using Ti02/CdSe and Sn02/CdSe
thin films [5.67-5.71]. Efforts have also been made to employ composite semiconduc­
tor systems in dye-sensitization systems [5.11, 5.72, 5.73].
An example of dye sensitization of an Sn02/CdS composite system is shown
in Figure 5.2. CdS is a short bandgap semiconductor (Eg = 2.5 eV) with a conduction

Fig. 5.2 Sensitization of Sn02-CdS coupled semiconductor with a dye (S) molecule.
148 Dye-Sensitized Solar Cells

band (-0.8 V vs NHE) that is more negative than that of Sn0 2 (0.0 V vs. NHE). Upon
excitation of the sensitizing dye, the electron is first injected into the CdS layer. The
photoinjected electrons are then quickly transferred from CdS to Sn0 2 nanocrystallites,
thus decreasing the probability of a back electron transfer process. The usefulness of
a composite semiconductor system in Ru(II) polypyridyl complex-based photochemi­
cal solar cells has been explored by several research groups [5.71, 5.73-5.75].
Pump-probe spectroscopy is a convenient way of monitoring both charge
injection and the charge recombination process, and for establishing the charge
rectification role of composite semiconductor systems. The transient spectrum that
was recorded 50 ns after the 532-nm laser pulse excitation of OTE/Sn02/Ru(II) and
OTE/Sn02/CdS/Ru(II) was significantly different (Fig. 5.3). (It should be noted that
532-nm laser pulse excitation ensured a selective excitation of Ru(II) since CdS and
Sn0 2 have negligible absorption at this excitation wavelength.) The lack of an absorb-
ance maximum at 380 nm indicated the decay of the excited state within the time
period of 50 ns. However, following the quenching of the excited state, a long-lived
bleaching at 460 nm persisted. This was indicative of the formation of the oxidation
product, Ru(III). The bleaching observed at 397 nm (a wavelength corresponding to
the isosbestic point of Ru(II)* and Ru(II) absorption) further confirmed the fact that
only the electron transfer product, Ru(III), contributed to the transient bleaching. A
similar observation of Ru(III) formation in a colloidal Sn0 2 suspension has also been
observed by our group [5.76] and by Ford and Rodgers [5.77-5.79].
If indeed the Sn02/CdS semiconductor composite system has a beneficial effect
by improving the charge separation efficiency, we should be able to see its influence
directly on the back electron transfer process. The rate constant of back electron trans­
fer obtained for an Sn02/CdS/Ru(II) film was smaller by a factor of 2-3 than the cor­
responding rate constant of an Sn02/Ru(II) film. This slower recovery of the transient
bleaching, observed at both monitoring wavelengths (460 nm), provided supportive

Fig. 5.3 Left: Transient spectra recorded 50 ns after laser pulse (532 nm) excitation of Ru(II)
modified semiconductor thin films coated on OTE: (a) OTE/Si02/Ru(II), (o); and (b) OTE/
Sn02/Ru(II), (·), and (c) OTE/Sn02/CdS/Ru(II), (D). Right: Absorption-time profiles at 460 nm
recorded following the 532 nm laser pulse excitation of OTE/Sn02/Ru(II), (o) and OTE/Sn02/
CdS/Ru(II), (·). Adapted from reference [5.74].
Photosensitization of Sn02 and other oxides 149

evidence of an improved charge separation in the Sn02/CdS composite films. Such an


enhancement in the charge rectification in Sn02/CdS composite film thus accounted
for the increased efficiency of the charge carrier accumulation.

5.3 SnO2-C60-Ru(bpy)f SYSTEM

Most of the visible excitation in dye-sensitized solar cells is centered around Ru(II)
inducing charge injection into semiconductor nanocrystallites. One interesting pos­
sibility is to coat an additional layer of C60 clusters so that it can be excited with
long-wavelength excitation and generate a triplet-excited C60, thus forming C6o ions
within the cluster framework. The C60/C6o redox couple acts as an electron relay to
regenerate the sensitizer. The present approach of using C60/C6o as an electron shuttle
paves the way towards improved photoelectrochemical cells [5.80].
An earlier study presented a photogalvanic mechanism of photocurrent genera­
tion at a C60 cluster-modified Sn0 2 electrode [5.81]. The excited C60 clusters inter­
acted with Γ to produce C60 anions in the cluster (reaction 5.1).

(5.1)

These formed C60 anions were relatively long-lived [5.82] and delivered
charge to the collecting electrode surface, thereby producing an anodic photocurrent.
The maximum IPCE observed for a C60 electrode alone is usually low (less than 5 %).
The mechanism of photocurrent generation using C60 clusters deposited on an Sn0 2
film and the iodide couple as the redox electrolyte can be found elsewhere [5.81].
If the C60 cluster film is deposited on an Ru(II) polypyridyl complex-modified
Sn0 2 film, it acts as an electron relay. On the other hand, the formation of C60 anions
producing a photocurrent, as well as a charge injection process, can cooperatively
contribute to the photocurrent generation, and the combined effect is smaller than the
enhanced IPCE observed with an OTE/SnO2/Ru(II)/(C60)n electrode. The maximum
IPCE obtained from spectrum d, which is an additive spectrum of a and b (Fig. 5.4),
is only marginally higher than that of the OTE/Sn02/Ru(II) electrode. These results
point out that the observed enhancement arose from the indirect participation of C60
cluster/C60 anion.
The photogenerated C6o (reaction 5.1) is electroactive and is long-lived
(>100/is) when generated in clusters and films [5.82]. It is capable of delivering
charge to oxides such as Sn0 2 [5.81]. Since the reduction potential of C60 is -0.2 V
vs. NHE, we expect a favorable electron transfer to Sn0 2 (ECB = 0 V vs. NHE), but
not to Ti0 2 (ECB = -0.5 V vs. NHE). This argument was supported by the fact that no
photocurrent generation was observed when Sn0 2 was replaced by Ti0 2 . The absence
of photocurrent for the OTE/TiO2/C60 electrode demonstrated that C6o ions cannot
directly transfer electrons to Ti0 2 .
In a dye-sensitized solar cell, iodide ions are capable of intercepting the oxi­
dized sensitizer (Ru(III)) and facilitate quick regeneration (reaction 5.2).

(5.2)
150 Dye-Sensitized Solar Cells

Fig. 5.4 The incident photon to charge carrier generation efficiency (IPCE) of modified nanos-
tructured Sn02 electrodes immersed in acetonitrile containing Γ and I~ electrolyte. Adapted
from reference [5.80].

The results presented in Figure 5.2 further point to the alternate pathway of
using C60 anions as the electron donors to adjacent Ru(III) species formed at the Ti0 2
interface (reaction 5.3).

(5.3)

Thus, the C60 layer shields most of the Ru(II)* from the triiodide couple while
at the same time acting as an electron mediator to regenerate the sensitizer.
The mechanism by which C60/C6o shuttles electrons and regenerates the sen­
sitizer is shown in Figure 5.5. The absorption of C60 clusters extends well into the IR
region and, when excited in the presence of iodide, C60 undergoes electron transfer
to generate C6o ions. The C6o ion becomes stabilized in the cluster and donates elec­
trons whenever it encounters an oxidized sensitizer (viz., Ru(III)). Neither the C60 nor
C6o has any significant influence on the excited state dynamics of Ru(IP). The only
role of the C60 cluster is to shuttle the electron between the Ru(III) and I~. A similar
mediating role of C60/C6o in the electrocatalytic oxidation of ferrocene has also been
demonstrated in a cyclic voltammetric study [5.83]. Moreover, charge transport by
lateral electron hopping within the fullerene monolayer, is observed by depositing
these clusters on insulating oxide (Zr0 2 ). Apparent diffusion coefficients as high as
1.5 x 10~8 cm2 s_1 were measured for the electron hopping process in this case [5.84].
The energy level diagram illustrating various charge transfer processes contributing
to the enhancement of the photocurrent generation of an OTE/Ti02/Ru(II)/(C6o)n elec­
trode is shown in Figure 5.5.
Photosensitization of Sn02 and other oxides 151

Fig. 5.5 Energy levels (not to scale) of various redox couples responsible for the charge trans­
fer processes. Adapted from reference [5.80].

5.4 PROBING THE INTERACTION OF AN EXCITED STATE


SENSITIZER WITH THE REDOX COUPLE

If we are interested in maximizing the photoconversion efficiency, it is essential to


minimize the excited state interaction with the redox couple. Here, we discuss a con­
cept of using C60 clusters as electron shuttles that effectively regenerate the sensitizer,
but at the same time minimize a direct interaction between the excited sensitizer and
the redox couple. The principle of this approach is illustrated in Figure 5.6.
The redox couple that has been found to be the most effective for regenerating
the sensitizer is the I~/l~ couple. The iodide ions donate electrons to the oxidized sensi­
tizer, thereby minimizing the loss of electrons in charge recombination [5.85-5.87]. A
high concentration of iodide is necessary to regenerate the sensitizer as it reduces the
Ru(III) with a rate constant of 1 x 1010 M_1 s_1. Although most of the charge injection
from the excited sensitizer into the semiconductor is completed in the subpicosecond
to nanosecond timescale [5.26, 5.35, 5.88], the presence of high concentrations of
I" and I3 increases the probability of their interaction with the excited state of the
sensitizer. Usually, such an energy loss due to quenching of the excited sensitizer by
the redox couple is thought to be small, although recent studies have demonstrated a
reductive quenching phenomenon [5.89, 5.90].
In an earlier study, we estimated the bimolecular quenching rate constant
between excited Ru(II) and Γ to be 1 x 108 s_1 [5.7]. We also measured the decay
rate constant of Ru(II)* on nanostructured silica film. The silica films were deposited
in the same manner as the Sn0 2 or Ti0 2 films. Since silica is an insulator, it does not
interact with the excited Ru(II). This rendered it possible to probe the quenching proc­
ess of Ru(II)* by I3. A solution of I~ was prepared by mixing equal concentrations
of I2 and Lil solution in acetonitrile. An acetonitrile solution with a known concen­
tration of I3 was then transferred to a square cuvette containing an OTE/Si02/Ru(II)
electrode. The electrode configuration was such that its backside was pressed against
152 Dye-Sensitized Solar Cells

Fig. 5.6 The use of an electron mediator to shuttle electrons for sensitizer regeneration. Adapted
from reference [5.80].

the cuvette and excited with a 337-nm laser pulse. Such a backside excitation of the
electrode minimized filtering of the excitation pulse by the I~ present in the solution.
Emission (600 nm) decay traces were collected in front-face geometry.
Figures 5.7(a) and (b) show emission decay at 600 nm of OTE/Si02/Ru(II) and OTE/
SiO2/Ru(II)/(C60)n at varying concentrations of I~. The excited Ru(II) bound to the
Si0 2 surface was readily quenched by I~. This was evidenced from the decreased
emission lifetime of Ru(II)* with an increasing concentration of I~. From the depend­
ence of the emission decay rate constant versus the I~ concentration, we obtained a
bimolecular rate constant of 1.9 x 1010 M_1 s_1. The fact that the rate constant was
close to the value of a diffusion-controlled reaction suggested the ability of I~ to effec­
tively intercept excited Ru(II)*. To the best of our knowledge, this is the first report
that directly probes the quenching between Ru(II)* and I~.
Ru(II)* can be quenched both reductively and oxidatively. In previous studies,
we have probed the reductive quenching of Ru(II)* by Γ by adsorbing the sensitizer
on a Si0 2 surface. The bimolecular quenching for this reaction was determined to
be 1 x 108 M_1 s_1 [5.7]. In the present investigation, the quenching of Ru(II)* by
I3 was oxidative with a greater rate constant than the one observed with iodide. The
undesirable quenching by the redox couple can pose a problem if it competes with the
charge injection from Ru(II)* to Sn0 2 or Ti0 2 . Considering the high concentration of
iodide and triiodide employed in the electrolyte, such a competitive process cannot
be ruled out.
The emission decay traces in Figure 5.7(b) showed little variation when the
concentration of I~ in solution was increased from 0 to 0.8 mM. The deposition of
C60 clusters on Ru(II) thus rendered it possible to minimize the interaction between
Ru(II)* and I~. This observation explains the beneficial role towards enhancing IPCE
of Sn02/Ru(II) and Ti02/Ru(II) films. Nevertheless, in a dye-sensitized solar cell,
Photosensitization of Sn02 and other oxides 153

Fig. 5.7 Emission decay traces of (a) OTE/Si02/Ru(II) and (b) (a) OTE/SiO2/Ru(II)/(C60)n
electrodes at varying concentrations of I~: (A) 0, (B) 0.2, (C) 0.4, and (D) 0.8 mM. Excitation
occurred at 337 nm and the emission was monitored at 600 nm. Adapted from reference [5.80].

where the charge injection is expected to be fast (in the subpicosecond - nanosecond
time scale) [5.26, 5.35, 5.88], the high quenching rate constant of I~ can be a limiting
factor, especially when one employs the redox couple at relatively high concentration
levels.

5.5 SENSITIZATION OF NANOTUBE ARRAYS

Recently, interest has been devoted to developing ordered arrays of Ti0 2 nanotubes
either by electrochemical etching of Ti foil in a fluoride medium or by depositing
Ti0 2 rods on a conducting surface [5.91-5.97]. Using this strategy, nanotube-[5.98]
and nanowire-based [5.99] DSCs have been reported. A dye-sensitized solar cell in
which the traditional nanoparticle film is replaced by a dense array of oriented, crys­
talline ZnO nanowires ensures the rapid collection of carriers generated throughout
the device [5.99]. Compared to mesoscopic semiconductor films, the ordered arrays
of tubes, wires and rods provide a well-defined architecture. An improvement in the
electron transport observed in a ZnO array has been attributed to the decrease in the
number of grain boundaries [5.100]. A ZnO rod-array employed in one such investi­
gation is shown in Figure 5.8.
Another approach involves the use of an SWCNT network on a conducting
electrode surface to promote charge transport in mesoscopic Ti0 2 films [5.102]. The
electrons injected from the excited dye into Ti0 2 nanoparticles are then transferred
through a SWCNT scaffold to generate a photocurrent. The semiconducting prop­
erty of SWCNTs has been successfully exploited to improve the performance of
organic photovoltaic cells [5.103] and fuel cells [5.104, 5.105]. While no net increase
in power conversion efficiency has been observed, an increase in photon-to-current
efficiency (IPCE) represents the beneficial role of SWCNTs as a conducting scaffold
to facilitate charge separation and charge transport in nanostructured semiconduc­
tor films. Such a nanowire/nanoparticle architecture is likely to play an important
154 Dye-Sensitized Solar Cells

Fig. 5.8 Left: A schematic representation of a ZnO-Nanorod array sensitized with zinc porphy-
rin. An FESEM image of a ZnO nanorod array prepared by MOCVD is also shown. (Images
kindly provided by Prof. Elena Galoppini, reference [5.100]) Right: FESEM images of a nano-
tube-array sample grown by electrochemical etching. (From reference [5.101]).

role in improving the efficiency of nanostructured solar energy conversion devices,


e.g., dye-sensitized solar cells, and quantum dot solar cells, as well as in water
photoelectrolysis.

5.6 CHARGE SEPARATION OF ORGANIC CLUSTERS


AT AN Sn0 2 ELECTRODE SURFACE

Dyes absorbing in the infrared region are important for extending the photoresponse of
solar cells. This is due to them harvesting low-energy photons that usually go untapped.
Of particular interest are squaraine and croconate dyes, which exhibit strong absorption
in the infrared [5.106-5.108]. These dye molecules possess a donor-acceptor-donor type
structure and their absorption in the IR region can be tuned by varying donor moieties.
Carbocyanine dyes represent another class of dyes absorbing in the red-infrared region.
An interesting aspect of these dye molecules is their ability to undergo intermolecular
interaction and produce H-type and J-type aggregates [5.109-5.117]. The excitation of
the H-aggregate produces excitons that are capable of undergoing charge separation
on the Ti0 2 and Sn0 2 surface. When dye molecules adsorbed onto the Ti0 2 films are
excited with visible-IR light, we observe an anodic current generation. Only the aggre­
gates contribute to the photocurrent generation, and since the conduction band energy of
Ti0 2 (-0.5 V vs. NHE) is more negative than the oxidation potential, the excited mono­
mer form of the squaraine or croconate dyes cannot directly participate in the charge
injection process. As a result of this energy mismatch, it is not possible to observe
sensitized photocurrent generation from a dye monomer. However, the dye aggregates
are capable of generating photocurrent as the excitons formed during the excitation of
aggregates undergo charge separation, thus contributing to the photocurrent [5.118].
The low IPCE of these dye films is proof that the net charge separation is poor.
Photosensitization of Sn02 and other oxides 155

Another class of molecules with importance in solar cell development is rep­


resented by conjugated polymers [5.119, 5.120]. Efficiencies up to 5 % have been
achieved with polythiophene-based organic solar cells [5.121, 5.122]. Oligomeric
units provide a convenient way of assessing the excited state properties and fac­
tors limiting light energy conversion of conjugated polymers. The conjugation of a
phenylene-ethynylene moiety can be extended by linking it to a polycyclic aromatic
hydrocarbon such as pyrene (Py-OPE) [5.123].
When cast as a film on the Sn02-modified conducting glass electrode (OTE/
Sn0 2 ), a small fraction of excitons formed in this oligomer dissociate to produce a
long-lived charge-separated state. The role of the Sn0 2 interface in promoting charge
separation is realized from the photoelectrochemical measurements. Figure 5.9(a) dis­
plays the photocurrent action spectrum of the OTE/Sn02/Py-OPE electrode presented
in terms of the incident photon to charge carrier generation efficiency (IPCE) versus
the wavelength of excitation. The OTE/Sn02/Py-OPE demonstrates a photocurrent
response at wavelengths below 480 nm with a maximum IPCE value around 6 %.
Comparisons with the absorption spectrum of the Py-OPE film show that the origin of
the photocurrent corresponds to the excitation of Py-OPE. Although we were unable
to harvest most of the visible photons, the extended ^-conjugation with the pyrene
moiety was useful for rendering the phenylene-ethynylene oligomers responsive to
visible light (up to 480 nm). The IPCE value was nearly one order of magnitude
greater than the one reported for Ti0 2 films with monolayer coverage [5.124].
Figure 5.9(b) shows the power characteristics of the photoelectrochemical cell
comprised of OTE/Sn02/Py-OPE. A stable photocurrent of 0.25 mA/cm2 and an open
circuit voltage of 300 mV were observed under visible light irradiation of 100 mW/
cm2. The overall power conversion efficiency of the cell was -0.05 %. As indicated
in our transient absorption studies, only a small fraction of the excitons were able
to dissociate at the Sn0 2 interface. Moreover, the electron transport across the Sn0 2
network remained a bottleneck for the charge transport.

Fig. 5.9 (a) (A) An absorption spectrum and (B) a photocurrent action spectrum of OTE/
Sn02/1. (b) I-V characteristics of OTE/Sn02/Py-OPE electrode under white light (A > 350 nm,
input power 100 mW/cm2) illumination. (Electrolyte: 0.5 M Nal and 0.02 mM I2 in water).
Adapted from reference [5.123].
156 Dye-Sensitized Solar Cells

5.7 CONCLUDING REMARKS

The field of dye-sensitized solar cells has seen a phenomenal growth during the last
two decades. Significant advances have been made in understanding the various
charge transfer and redox processes in dye-sensitized solar cells. Recent approaches
to the design of photoactive nanostmctured architectures and inorganic-organic hybrid
assemblies offer new opprounities to further improve the performance of photochemi­
cal solar cells. Many of the newly synthesized sensitizer molecules and 1-D nano-
materials show potential for the development of economically viable and practically
appealing solar cells.

5.8 ACKNOWLEDGEMENTS

The author acknowledges the support of the office of Basic Energy Sciences of the
U.S. Department of Energy.

5.9 REFERENCES

[5.1] Kamat, P. V., Meeting the Clean Energy Demand: Nanostructure Architectures for Solar Energy
Conversion, J. Phys. Chem. C, 2007, 777, 2834-2860.
[5.2] Kamat, P. V., Quantum Dot Solar Cells. Semiconductor Nanocrystals as Light Harvesters,
J. Phys. Chem. C, 2008, 772, 18737-18753.
[5.3] Galoppini, E., Linkers for anchoring sensitizers to semiconductor nanoparticles, Coord.Chem
Rev, 2004, 248, 1283-1297.
[5.4] Grätzel, M., Solar Energy Conversion by Dye-Sensitized Photovoltaic Cells, Inorg. Chem., 2005,
44,6841-6851.
[5.5] Bisquert, J., Cahen, D., Hodes, G., Riihle, S., Zaban, A., Physical Chemical Principles of
Photovoltaic Conversion with Nanop articulate, Mesoporous Dye-Sensitized Solar Cells, J. Phys.
Chem. B, 2004,108, 8106-8118.
[5.6] Meyer, G. J., Molecular Approaches to Solar Energy Conversion with Coordination Compounds
Anchored to Semiconductor Surfaces, Inorg. Chem., 2005, 44, 6852-6864.
[5.7] Nasr, C., Hotchandani, S., Kamat, P. V., Role of Iodide in Photoelectrochemical Solar Cells.
Electron Transfer between Iodide Ions and Ruthenium Polypyridyl Complex Anchored on
Nanocrystalline Si02 and Sn02 Films, J. Phys. Chem. Β, 1998,102, 4944-4951.
[5.8] Kamat, P. V., Chauvet, J. P., Fessenden, R. W., Photoelectrochemistry in paniculate systems. 4.
Photosensitization of a Ti02 semiconductor with a chlorophyll analogue, J. Phys. Chem., 1986,
90, 1389-1394.
[5.9] Bedja, I., Kamat, P. V., Hotchandani, S., Fluorescence and photoelectrochemical behavior of
chlorophyll a adsorbed on a nanocrystalline Sn02film, J. Appl. Phys., 1996, 80, 4637-4643.
[5.10] Bedja, I., Hotchandani, S., Carpentier, R., Fessenden, R. W., Kamat, P. V., Chlorophyll b modified
nanocrystalline Sn02 semiconductor thin film as a photosensitive electrode, J. Appl. Phys., 1994,
75, 5444-5456.
[5.11] Hotchandani, S., Kamat, P. V., Modification of electrode surface with semiconductor colloids and
its sensitization with chlorophyll a, Chem. Phys. Lett, 1992, 797, 320-326.
[5.12] Hotchandani, S., Das, S., Thomas, K. G., George, M. V., Kamat, P. V., Interaction of semicon­
ductor colloids with J-aggregates of squaraine dye and its role in sensitizing nanocrystalline
semiconductor films, Res. Chem. Intermed., 1994, 20, 927-938.
Photosensitization of S n 0 2 and other oxides 157

[5.13] Kim, Y.-S., Liang, K., Law, K.-Y., Whitten, D. G., An investigation of photocurrent genera­
tion by squaraine aggregates in monolayer-modified Sn02 electrodes, J. Phys. Chem., 1994, 98,
984-988.
[5.14] Nasr, C , Liu, D., Hotchandani, S., Kamat, P. V., Dye capped semiconductor colloids. Excited
state and photosensitization aspects of Rhodamine 6G-H aggregates electrostatically bound to
Si02 andSn02 Colloids., J. Phys. Chem., 1996,100, 11054-11061.
[5.15] Ferrere, S., Zaban, A., Gregg, B. A., Dye sensitization of nanocrystalline tin oxide by perylene
derivative, J. Phys. Chem. B, 1997,101, 4490-4493.
[5.16] Liu, D., Kamat, P. V., Electrochemically active nanocrystalline Sn02 films. Surface modification
with thiazine and oxazine dye aggregates., J. Electrochem. Soc, 1995, 742, 835-839.
[5.17] Sykora, M., Petruska, M. A., Alstrum-Acevedo, J., Bezel, I., Meyer, T. J., Klimov, V. I.,
Photoinduced Charge Transfer between CdSe Nanocrystal Quantum Dots and Ru-Polypyridine
Complexes, J. Am. Chem. Soc, 2006,128, 9984-9985.
[5.18] Genscher, H., Willig, F., Reaction of excited dye molecules at electrodes, Top. Curr. Chem., 1976,
61, 31-84.
[5.19] Ryan, M. A., Spitler, M. T., Photoelectrochemistry and photochemistry of dyes adsorbed at semi­
conductor surfaces, J. Imaging Sei, 1989, 33, 46-49.
[5.20] Hashimoto, K., Hiramoto, M., Kajiwara, T., Sakata, T., Luminescence decays and spectra of
Ru(bpy)32+ adsorbed on Ti02 in vacuo and in the presence of water vapor, J. Phys. Chem., 1988,
92, 4636-4640.
[5.21] Hashimoto, K., Hiramoto, M., Lever, A. B. P., Sakata, T., Luminescence decay of'ruthenium(II)
complexes adsorbed on metal oxide powders in vacuo: Energy gap dependence of the electron-
transfer rate, J. Phys. Chem., 1988, 92, 1016-1018.
[5.22] Tani, T., New aspects of the electron transfer mechanism for spectral sensitization and supersen-
sitization., J. Imag. Sei., 1990, 34, 143-148.
[5.23] Tani, T., Suzumoto, T., Ohzeki, K., Energy gap dependence of efficiency of photoinduced elec­
tron transfer from cyanine dyes to silver bromide microcrystals in spectral sensitization, J. Phys.
Chem., 1990, 94, 1298-1300.
[5.24] Ryan, M. A., Fitzgerald, E. C., Spitler, M. T., Internal reflection flash photolysis study of
the photochemistry of eosin at Ti02 semiconductor electrodes, J. Phys. Chem., 1989, 93,
6150-6156.
[5.25] Sonntag, L. P., Spitler, M. T., Examination of the energetic threshold for dye-sensitized photocur­
rent at SrTiO3 electrodes, J. Phys. Chem., 1985, 89, 1453-1457.
[5.26] Fessenden, R. W., Kamat, P. V., Rate constants for charge injection from excited sensitizer into
Sn02, ZnO, and Ti02 semiconductor nanocrystallites, J. Phys. Chem., 1995, 99, 12902-12906.
[5.27] Rehm, J. M., McLendon, G. L., Nagasawa, Y, Yoshihara, K., Moser, J., Graetzel, M.,
Femtosecond Electron-Transfer Dynamics at a Sensitizing Dye-Semiconductor (Ή02) Interface,
J. Phys. Chem., 1996,100, 9577-9578.
[5.28] Burfeindt, B., Hannappel, T., Storck, W, Willig, F., Measurement of Temperature-Independent
Femtosecond Interfacial Electron Transfer from an Anchored Molecular Electron Donor to a
Semiconductor as Acceptor, J. Phys. Chem., 1996,100, 16463-16465.
[5.29] Martini, I., Hartland, G., Kamat, P. V., Ultrafast Investigation of the Photophysics of Cresyl
Violet adsorbed onto Nanometer Sized Particles of Sn02 and Si02., J. Phys. Chem., 1997,101B,
4826-4830.
[5.30] Tachibana, Y, Moser, J. E., Graetzel, M., Klug, D. R., Durrant, J. R., Subpicosecond interfacial
charge separation in dye-sensitized nanocrystalline titanium dioxide films, J. Phys. Chem., 1996,
100, 20056-20062.
[5.31] Martini, I., Hodak, J., Hartland, G., Kamat, P. V., Ultrafast study of interfacial electron trans­
fer between 9-anthracene-carboxylate and Ti02 semiconductor particles., J. Chem. Phys., 1997,
107, 8064-8072.
[5.32] Hannappel, T., Burfeindet, B., Storck, W., Willig, F., Measurement of ultrafast photoinduced
electron transfer from chemically anchored Ru-dye molecules into empty electronic states in a
colloidal anatase Ti02film, J. Phys. Chem. B, 1997,101, 6799-6802.
[5.33] Randy, J., Ellingson, R. J., Asbury, J. B., Ferrere, S., Ghosh, H. N., Sprague, J. R., Lian, T., Nozik,
A. J., Dynamics of Electron Injection in Nanocrystalline Titanium Dioxide Films Sensitized with
158 Dye-Sensitized Solar Cells

[5.Ru(4,4'-dicarboxy-2,2'-bipyridine)2 (NCS)2] by Infrared Transient Absorption, J. Phys. Chem.


B, 1998, 702, 6455-6458.
[5.34] Asbury, J. B., Wang, Y. Q., Lian, T., Multiple-Exponential Electron Injection in Ru(dcbpy)2(SCN)2
Sensitized ZnO Nanocrystalline Thin Films, J. Phys. Chem. B, 1999,103, 6643-6647.
[5.35] Asbury, J. B., Randy, J., Ellingson, R. J., Ghosh, H. N., Ferrere, S., Nozik, A. J., Lian, T.,
Femtosecond IR Study of Excited-State Relaxation and Electron-Injection Dynamics of
Ru(dcbpy)2(NCS)2 in Solution and on Nanocrystalline Ti02 andAl203, J. Phys. Chem. B, 1999,
705,3110-3119.
[5.36] Asbury, J. B., Hao, E., Wang, Y, Lian, T., Bridge Length-Dependent Ultrafast Electron Transfer
from Re Polypyridyl Complexes to NanocrysallineTi02 Thin Films Studied by Femtosecond
Infrared Spectroscopy, J. Phys. Chem. B, 2000,104, 11957-11964.
[5.37] Bauer, C., Boschloo, G., Mukhtar, E., Hagfeldt, A., Electron Injection and Recombination in
Ru(dcbpy)2(NCS)2 Sensitized Nanostructured ZnO, J. Phys. Chem. B, 2001,105, 5585-5588.
[5.38] Kallioinen, J., Benkö, G., Sundström, V., Korppi-Tommola, J. E. I., Yartsev, A. R, Electron
Transfer from the Singlet and Triplet Excited States of Ru(dcbpy)2(NCS)2 into Nanocrystalline
Ή02 Thin Films, J. Phys. Chem. B, 2002,106, 4396-4404.
[5.39] Furube, A., Katoh, R., Hara, K., Murata, S., Arakawa, H., Tachiya, M., Ultrafast Stepwise Electron
Injection from Photoexcited Ru-Complex into Nanocrystalline ZnO Film via Intermediates at the
Surface, J. Phys. Chem. B, 2003, 707, 4162-4166.
[5.40] Horiuchi, H., Katoh, R., Hara, K., Yanagida, M., Murata, S., Arakawa, H., Tachiya, M., Electron
Injection Efficiency from Excited N3 into Nanocrystalline ZnO Films: Effect of (N3-Zn2+)
Aggregate Formation, J. Phys. Chem. B, 2003, 707, 2570-2574.
[5.41] Marcus, R. A., Sutin, N., Electron Transfers in Chemistry and Biology, Biochim. Biophys. Acta,
1985, 877, 265.
[5.42] Marcus, R. A., On Theory of Electron-Transfer Reactions. 6. Unified Treatmentfor Homogeneous
and Electrode Reactions, J. Chem. Phys., 1965, 43, 679-701.
[5.43] Gaal, D. A., Hupp, J. T., Thermally activated, inverted interfacial electron transfer kinetics: High
driving force reactions between tin oxide nanoparticles and electrostatically-bound molecular
reactants, J. Am. Chem. Soc, 2000, 722, 10956-10963.
[5.44] Kuciauskas, D., Freund, M. S., Gray, H. B., Winkler, J. R., Lewis, N. S., Electron Transfer
Dynamics in Nanocrystalline Titanium Dioxide Solar Cells Sensitized with Ruthenium or Osmium
Polypyridyl Complexes, J. Phys. Chem. B, 2001, 705, 392-403.
[5.45] Clifford, J. N., Palomares, E., Nazeeruddin, M. K., Gratzel, M., Nelson, J., Li, X., Long,
N. J., Durrant, J. R., Molecular Control of Recombination Dynamics in Dye-Sensitized
Nanocrystalline Ti02 Films: Free Energy vs Distance Dependence, J. Am. Chem. Soc, 2004,
126, 5225-5233.
[5.46] Kamat, R V., Bedja, I., Hotchandani, S., Patterson, L. K., Photosensitization of nanocrystalline
semiconductor films. Modulation of electron transfer between excited ruthenium complex and
Sn02 nanocrystallites with an externally applied bias., J. Phys. Chem., 1996, 700, 4900-4908.
[5.47] O'Regan, B., Graetzel, M., Fitzmaurice, D., Optical Electrochemistry I: Steady-state spectros­
copy of conduction band electrons in a metal oxide semiconductor electrode., Chem. Phys. Lett.,
1991,183, 89-93.
[5.48] O'Regan, B., Graetzel, M., Fitzmaurice, D., Optical Electrochemistry. 2. Real-time spectroscopy
of conduction band electrons in a metal oxide semiconductor electrode., J. Phys. Chem., 1991,
95, 10525-10528.
[5.49] Redmond, G., Fitzmaurice, D., Graetzel, M., Effect of surface chelation on the energy of an
intrab and surface state of a nanocrystalline Ti02film., J. Phys. Chem., 1993, 97, 6951-6954.
[5.50] Redmond, G., Fitzmaurice, D., Spectroscopic determination of fiatband potentials for polycrys-
talline Ti02 electrodes in nonaqueous solvents., J. Phys. Chem., 1993, 97, 1426-1430.
[5.51] Bedja, I., Hotchandani, S., Kamat, P. V., Photoelectrochemistry of quantized W03 colloids.
Electron storage, electrochromic, and photoelectrochromic effects, J. Phys. Chem., 1993, 97,
11064-11070.
[5.52] Hotchandani, S., Bedja, I., Fessenden, R. W., Kamat, P. V., Electrochromic and photoelectrochro­
mic behavior of thin WO3 films prepared from quantum size colloidal particles, Langmuir, 1994,
70, 17-22.
Photosensitization of S n 0 2 and other oxides 159

[5.53] Hagfeldt, A., Vlachopoulos, N., Graetzel, M., Fast electrochromic switching with nanocrystalline
oxide semiconductor films, J. Electrochem. Soc, 1994, 747, L82-L87.
[5.54] Breddels, P. A., Blasse, G., The luminescence properties of some ruthenium(II) complexes and
solid dye films in relation to their application in photo electro chemical cells., Ber Bunsenges.
Phys. Chem., 1982, 86, 676-680.
[5.55] Kay, A., Humphry-Baker, R., Graetzel, M., Artificial Photosynthesis. 2. Investigations on the
mechanism of Photosensitization ofNanocrystalline Ti02 Solar Cells by Chlorophyll Derivatives,
J. Phys. Chem., 1994, 98, 952-959.
[5.56] Goff, A. H., Falaras, P., Origin of new bands in the Raman spectra of dye monolayers adsorbed
on nanocrystalline Ti02., J. Electrochem. Soc, 1995, 742, L38-L41.
[5.57] Bedja, I., Hotchandani, S., Kamat, P. V., Transient absorption spectroscopy of nanostruc-
tured films. An in situ spectroelectrochemical investigation of photosensitization process., J.
Electroanal. Chem., 1996, 407, 237-241.
[5.58] O'Regan, B., Moser, J., Anderson, M., Graetzel, M., Vectorial electron injection into transpar­
ent semiconductor membranes and electric field effects on the dynamics of light-induced charge
separation, J. Phys. Chem., 1990, 94, 8720-8726.
[5.59] Hotchandani, S., Kamat, P. V., Charge-transfer processes in coupled semiconductor systems.
Photochemistry and photoelectrochemistry of the colloidal CdS-ZnO system, J. Phys. Chem.,
1992, 96, 6834-6839.
[5.60] Vinodgopal, K., Kamat, P. V., Enhanced rates of photo catalytic degradation of an azo dye using
Sn02/Ti02 coupled semiconductor thin films., Environ. Sei. Technol, 1995, 29, 841-845.
[5.61] Vinodgopal, K., Bedja, I., Kamat, P. V., Nano structured semiconductor films for photo catalysis.
Photoelectrochemical behavior ofSn02/Ti02 coupled systems and its role in photocatalytic deg­
radation of a textile azo dye., Chem. Mater, 1996, 8, 2180-2187.
[5.62] Spanhel, L., Weiler, H., Henglein, A., Photochemistry of semiconductor colloids. 22. Electron
injection from illuminated CdS into attached Ti02 and ZnO particles, J. Am. Chem. Soc, 1987,
109, 6632-6635.
[5.63] Gopidas, K. R., Bohorquez, M., Kamat, P. V., Photoelectrochemistry in semiconductor paniculate
systems. 16. Photophysical and photochemical aspects of coupled semiconductors. Charge-trans­
fer processes in colloidal CdS-Ti02 and CdS-Agi systems, J. Phys. Chem., 1990, 94, 6435-6440.
[5.64] Kennedy, R., Martini, I., Hartland, G., Kamat, P. V., Capped Semiconductor Colloids: Synthesis
and Photochemistry of CdS Capped Sn02 Nanocrystallites, Proc. Indian Acad. Sei. (Chem. Sei.),
1997,109, 497-507.
[5.65] Rabani, J., Sandwich colloids of ZnO and ZnS in aqueous solutions, J. Phys. Chem., 1989, 93,
7707-7713.
[5.66] Kamat, P. V., Patrick, B., Photochemistry and photophysics of ZnO colloids, in Symp. Electron.
Ionic Prop. Silver Halides. 1991. Springfield, Va: The Society for Imaging Science and
Technology.
[5.67] Liu, D., Kamat, P. V, Electrochemical rectification in CdSe + Ti0 2 coupled semiconductor films,
/. Electroanal. Chem. Interfacial Electrochem, 1993, 347, 451-456.
[5.68] Liu, D., Kamat, P. V, Photoelectrochemical behavior of thin CdSe and coupled Ti02/CdSe semi­
conductor films, J. Phys. Chem., 1993, 97, 10769-10773.
[5.69] Vogel, R., Hoyer, P., Weller, H., Quantum-sized PbS, CdS, Ag2S, Sb2S3 and Bi2S3 particles as
sensitizers for various nanoporous wide-bandgap semiconductors, J. Phys. Chem., 1994, 98,
3183-3188.
[5.70] Matsumoto, H., Matsunaga, T., Sakata, T., Mori, H., Yoneyama, H., Size dependent fluores­
cence quenching of CdS nanocrystals caused by Ti02 colloids as a potential variable quencher,
Langmuir, 1995, 77, 4283-4287.
[5.71] Nasr, C., Kamat, P. V, Hotchandani, S., Photoelectrochemical behavior of coupled Sn02/CdSe
nanocrystalline semiconductor films., J. Electroanal. Chem., 1997, 420, 201-207.
[5.72] Chappel, S., Chen, S.-G., Zaban, A., Ti02-Coated Nanoporous Sn02 Electrodes for Dye-
Sensitized Solar Cells, Langmuir, 2002,18, 3336-3342.
[5.73] Kay, A., Grätzel, M., Dye-Sensitized Core-Shell Nanocrystals: Improved Efficiency ofMesoporous
Tin Oxide Electrodes coated with a Thin Layer of an Insulating Oxide, Chem. Mater, 2002, 74,
2930-2935.
160 Dye-Sensitized Solar Cells

[5.74] Nasr, C , Hotchandani, S., Kim, W. Y., Schmehl, R. H., Kamat, R V., Photoelectrochemistry of
composite semiconductor thin films. Photosensitization of Sn02/CdS coupled nanocrystallites
with a Ruthenium complex, J. Phys. Chem. B, 1997,101, 7480-7487.
[5.75] Nasr, C , Hotchandani, S., Kamat, P. V., Photoelectrochemistry of composite semiconductor thin
films. II. Photosensitization of Sn02/Ti02 coupled system with a ruthenium polypyridyl complex.,
J. Phys. Chem. B, 1998,102, 10047-10056.
[5.76] Liu, D., Hug, G. L., Kamat, P. V., Photochemistry on Surfaces. Intermolecular energy and elec­
tron transfer processes between excited Ru{bpy)]+ and H- aggregates ofCresyl Violet on Si02 and
Sn02 colloids, J. Phys. Chem., 1995, 99, 16768-16775.
[5.77] Ford, W. E., Rodgers, M. A. J., Interfacial electron transfer in colloidal Sn02 hydrosols photo­
sensitized by electrostatically and covalently attached Ru(II) polypyridine complexes, J. Phys.
Chem., 1994, 98,3822-3831.
[5.78] Ford, W. E., Rodgers, M. A. J., Photosensitization of colloidal Sn02 in an asymmetric supported
bilay er composed of surfactants., J. Phys. Chem., 1995, 99, 5139-5145.
[5.79] Ford, W. E., Rodgers, M. A. J., Kinetics ofNitroxyl Radical Oxidation by Ru(bpy)33+ following
Photosensitization of Antimony-Doped Tin Dioxide Colloidal Particles, J. Phys. Chem., 1997,
101, 930-936.
[5.80] Kamat, P. V., Haria, M., Hotchandani, S., C60 Cluster as an Electron Shuttle in a Ru(II)-Polypyridyl
Sensitizer Based Photochemical Solar Cell, J. Phys. Chem. B, 2004,108, 5166-5170.
[5.81] Kamat, P. V., Barazzouk, S., Thomas, K. G., Hotchandani, S., Electrodeposition of C60 Clusters
on Nanostructured Sn02 Films for Enhanced Photocurrent Generation, J. Phys. Chem. B, 2000,
104, 4014-4017.
[5.82] Thomas, K. G., Biju, V., George, M. V., Guldi, D. M., Kamat, P. V., Photoinduced Charge
Separation and Stabilization in Clusters of a Fullerene-Aniline Dyad, J. Phys. Chem. B, 1999,
103, 8864-8869.
[5.83] Barazzouk, S., Hotchandani, S., Kamat, P. V., Unusual Electrocatalytic Behavior of Ferrocene
Bound Fullerene Cluster Films, J. Mater. Chem., 2002,12, 2021-2025.
[5.84] Papageorgiou, N., Grätzel, M., Enger, O., Bonifazi, D., Diederich, F., Lateral Electron Transport
inside a Monolayer of Derivatized Fullerene s Anchored on Nanocrystalline Metal Oxide Films,
J. Phys. Chem. B, 2002,106, 3813-3822.
[5.85] Dressick, W. J., Meyer, T. J., Durham, B., Rillema, D. P., Excited state photoelectrochemical cells
for the generation ofH2 and 02 based on Ru(bpy)23+, Inorg. Chem., 1982, 21, 3451-3458.
[5.86] Duffy, N. W., Peter, L. M., Rajapakse, R. M. G., Wijayantha, K. J. U., Investigation of the Kinetics
of the Back Reaction of Electrons with Tri-Iodide in Dye-Sensitized Nanocrystalline Photovoltaic
Cells, J. Phys. Chem. B, 2000,104, 8916-8919.
[5.87] Heimer, T. A., Heilweil, E. J., Bignozzi, C. A., Meyer, G. J., Electron Injection, Recombination,
and Halide Oxidation Dynamics at Dye-Sensitized Metal Oxide Interfaces, J. Phys. Chem. A.,
2000,104, 4256-4262.
[5.88] Ramakrishna, S., Willig, F., Pump-Probe Spectroscopy of Ultrafast Electron Injection from the
Excited State of an Anchored Chromophore to a Semiconductor Surface in UHV: A Theoretical
Model, J. Phys. Chem. B, 2000,104, 68-77.
[5.89] Thompson, D. W., Kelly, C. A., Farzad, F., Meyer, G. J., Sensitization of Nanocrystalline
Ti02 Initiated by Reductive Quenching of Molecular Excited States, Langmuir, 1999, 75,
650-653.
[5.90] Bergeron, B. V., Meyer, G. J., Reductive electron transfer quenching of MLCT excited states
bound to nanostructured metal oxide thin films, J. Phys. Chem. B, 2003,107, 245-254.
[5.91] Imai, H., Takei, Y, Shimizu, K., Matsuda, M., Hirashima, H., Direct preparation of anatase Ti02
nanotubes in porous alumina membranes, J. Mater. Chem., 1999, 9, 2971-2972.
[5.92] Tian, Z. R. R., Voigt, J. A., Liu, J., McKenzie, B., Xu, H. F , Large oriented arrays and continu­
ous films of Ti02-based nanotubes, J. Am. Chem. Soc, 2003,125, 12384-12385.
[5.93] Wu, J. J., Yu, C. C., Aligned Ti02 nanorods and nanowalls, J. Phys. Chem. B, 2004, 108,
3377-3379.
[5.94] Ruan, C., Paulose, M., Varghese, O. K., Mor, G. K., Grimes, C. A., Fabrication of Highly
Ordered Ti02 Nanotube Arrays Using an Organic Electrolyte, J Phys. Chem. B, 2005, 109,
15754-15759.
Photosensitization of S n 0 2 and other oxides 161

[5.95] Mor, G. K., Varghese, O. K., Paulose, M., Shankar, K., Grimes, C. A., A review on highly ordered,
vertically oriented Ti02 nanotube arrays: Fabrication, material properties, and solar energy
applications, Sol. Energy Mater Solar Cells, 2006, 90, 2011-2075.
[5.96] Paulose, M., Shankar, K., Varghese, O. K., Mor, G. K., Grimes, C. A., Application of highly -
ordered Ti02 nanotube-arrays in heterojunction dye-sensitized solar cells, Journal of Physics
D-Applied Physics, 2006, 39, 2498-2503.
[5.97] Adachi, M., Murata, Y, Okada, I., Yoshikawa, S., Formation of titania nanotubes and applica­
tions for dye-sensitized solar cells, J. Electrochem. Soc, 2003,150, G488-G493.
[5.98] Macak, J. M., Tsuchiya, H., Ghicov, A., Schmuki, P., Dye-sensitized anodic Ti02 nanotubes,
Electrochem. Commun., 2005, 7, 1133-1137.
[5.99] Law, M., Greene, L. E., Johnson, J. C., Saykally, R., Yang, P., Nanowire dye-sensitized solar
cells, Nature Mater, 2005, 4, 455-459.
[5.100] Galoppini, E., Rochford, J., Chen, H., Saraf, G., Lu, Y, Hagfeldt, A., Boschloo, G., Fast Electron
Transport in Metal Organic Vapor Deposition Grown Dye-sensitized ZnO Nanorod Solar Cells,
J. Phys. Chem. B, 2006, 16159-16161.
[5.101] Paulose, M., Shankar, K., Yoriya, S., Prakasam, H. E., Varghese, O. K., Mor, G. K., Latempa, T.
A., Fitzgerald, A., Grimes, C. A., Anodic Growth of Highly Ordered Ti02 Nanotube Arrays to 134
ßm in length, J. Phys. Chem. B, 2006,110, 16179-16184.
[5.102] Brown, P. R., Takechi, K., Kamat, P. V, Single-Walled Carbon Nanotube Scaffolds for Dye-
Sensitized Solar Cells, J. Phys. Chem. C, 2008, 772, 4776-4782.
[5.103] Hoppe, H., Sariciftci, N. S., Organic solar cells: An overview J. Mater. Res., 2004, 19,
1924-1945.
[5.104] Girishkumar, G., Hall, T. D., Vinodgopal, K., Kamat, P. V, Single Wall Carbon Nanotube
Supports for Portable Direct Methanol Fuel Cells, J. Phys. Chem. B, 2006, 770, 107-114.
[5.105] Kongkanand, A., Vinodgopal, K., Kuwabata, S., Kamat, P. V, Highly-dispersed Pt catalysts on
Single-Walled Carbon Nanotubes and Their Role in Methanol Oxidation, J. Phys. Chem. B, 2006,
770, 16185-16192.
[5.106] Law, K. Y, Organic Photoconductive Materials: Recent trends and developments, Chem. Rev.,
1993, 93, 449-486.
[5.107] Jürgen Fabian, R. Z., The Search for Highly Colored Organic Compounds, Angew. Chem., Int.
Ed., 1989, 28, 611-69A.
[5.108] Fabian, J., Nakazumi, H., Matsuoaka, M., Near-Infrared Absorbing Dyes, Chem. Rev., 1992, 92,
1197-1226.
[5.109] Serpone, N., Sahyun, M. R. V, Photophysics of Dithiacabocyanine Dyes:Subnanosecond relaxa­
tion dynamics of a dithia-2,2' -carbocyanine dye and its 9-methyl-substituted me so analog, J.
Phys. Chem., 1994, 98, 734-737.
[5.110] Soper, S. A., Mattingly, Q. L., Steady-state and picosecond fluorescence studies of non-radiative
pathways in tricarbocyanine dyes: Implications to the design of near-IR Fluorochromes with
high fluorescence efficiencies, J. Am. Chem. Soc, 1994,116, 3744-3752.
[5.111] Chen, S.-Y, Horng, M.-L., Quitevis, E. L., Picosecond spectroscopic studies of electronic energy
relaxation in H-aggregates of 1 ,T'-diethyl-2,2''-dicarbocyanine on colloidal silica., J. Phys.
Chem., 1989, 93, 3683-3688.
[5.112] Tani, T., Suzumoto, T., Kemnitz, K., Yoshihara, K., Picosecond kinetics of light induced electron
transfer from J-aggregated cyanine dyes to AgBr crystals, J. Phys. Chem., 1992, 96, 2778-2783.
[5.113] Noukakis, D., Van der Auweraer, M., Toppet, S., De Schryver, F. C , Photophysics of a thiacar-
bocyanine dye in organic solvents., J. Phys. Chem., 1995, 99, 11860-11866.
[5.114] Chibisov, A. K., Zakharova, G. V, Goerner, H., Sogulyaev, Y A., Mushkalo, I. L., Tolmachev, A.
I., Photorelaxation processes in covalently linked indocarbocyanine and thiacarbocyanine dyes,
J. Phys. Chem., 1995, 99, 886-93.
[5.115] Trosken, B., Wiilig, F., Schwarzburg, K., Ehret, A., Spitler, M., Electron transfer quenching of
excited J-aggregate dyes on AgBr microcrystals between 300 and 5K, J. Phys. Chem., 1995, 99,
5152-5160.
[5.116] Yonezawa, Y, Ishizawa, H., Luminescence properties of the mixed J-aggregate consisting of
two kinds of thiacarbocyanine dyes or naphthothiacarbocyanine dyes having meso-substituent,
J. Luminescence, 1996, 69, 141-150.
162 Dye-Sensitized Solar Cells

[5.117] Amn, K. T., Epe, B., Ramaiah, D., Aggregation Behavior of halogenated squaraine dyes in buffer,
electrolytes, organized media, and DNA, J. Phys. Chem. B, 2002,106, 11622-11627.
[5.118] Gregg, B. A., Excitonic Solar Cells, J. Phys. Chem. B, 2003,107, 4688-4698.
[5.119] Silverman, E. E., Cardolaccia, T., Zhao, X. M., Kim, K. Y., Haskins-Glusac, K., Schanze, K. S.,
The triplet state in Pt-acetylide oligomers, polymers and copolymers, Coord. Chem Rev, 2005,
249, 1491-1500.
[5.120] Pingree, L. S. C., MacLeod, B. A., Ginger, D. S., The changing face of PEDOT : PSS films:
Substrate, bias, and processing effects on vertical charge transport, J. Phys. Chem. C, 2008, 772,
7922-7927.
[5.121] Campoy-Quiles, M., Ferenczi, T., Agostinelli, T., Etchegoin, P. G., Kim, Y, Anthopoulos, T.
D., Stavrinou, P. N., Bradley, D. D. C , Nelson, J., Morphology evolution via self-organization
and lateral and vertical diffusion in polymer: fullerene solar cell blends, Nature Mater, 2008, 7,
158-164.
[5.122] Quist, P. A. C , Sweelssen, J., Koetse, M. M., Savenije, T. J., Siebbeles, L. D. A., Formation and
decay of charge carriers in bulk heterojunctions of MDMO-PPV or P3HTwith new n-type con­
jugated polymers, J. Phys. Chem. C, 2007, 777, 4452-4457.
[5.123] Matsunaga, Y, Takechi, K., Akasaka, T., Ramesh, A. R., James, P. V., Thomas, K. G., Kamat,
P. V., Excited State and Photoelectrochemical Behavior of Pyrene Linked Phenyleneethynylene
Oligomer, J. Phys. Chem. B, 2008, 772, 14539-14547.
[5.124] Taratula, O., Rochford, J., Piotrowiak, P., Galoppini, E., Carlisle, R. A., Meyer, G. J., Pyrene-
terminated phenylenethynylene rigid linkers anchored to metal oxide nanoparticles, J. Phys.
Chem. B, 2006, 770, 15734-15741.
CHAPTER 6

SOLID-STATE DYE-SENSITIZED SOLAR


CELLS INCORPORATING MOLECULAR
HOLE-TRANSPORTERS
Henry J. Snaith

6.1 INTRODUCTION

In the current economic, environmental and social climate, creating a revolutionary


low-cost photovoltaic system suitable for large scale power generation is of utmost
urgency. Dye-sensitized solar cells (DSCs) are receiving considerable academic and
industrial attention for this purpose, since they promise to convert solar to electrical
energy at a fraction of the cost of traditional semiconductor-based photovoltaics. As
has been discussed in detail in previous chapters, the state-of-the-art DSCs are electro­
chemical solar cells, incorporating an iodide/triiodide redox couple, and have verified
solar-to-electrical power conversion efficiencies in excess of 11 %. Although initial
steps to commercialization have been taken, it is industrially much more attractive to
remove the liquid phase of the device and realize an efficient all solid-state version
of the dye-sensitized solar cell. One only has to look at the short lifetime and leakage
issue with disposable batteries to understand that this is a real issue, especially if low-
cost sealing is required. Substantial research efforts involve using gel and solid-state
electrolytes. These show great promise and are likely to be a key component of the first
commercially viable DSC concepts. An alternative, which furthermore negates the cor­
rosive nature of the electrolyte, is to use a solid-state hole conductor. Amongst the pos­
sible inorganic p-type semiconductors, copper-based compounds appear to be suitable
alternatives. These hole-transporters can be cast from solution or vacuum deposition to
fully interpenetrate the mesoporous Ti0 2 . A solid-state device based on CuI was first
demonstrated by Tennakone et al. [6.1], and a power conversion efficiency as high as
4.7 % has been reported [6.2]. Though CuI is a very promising material, the cells are
generally unstable, which is attributed to a stoichiometric excess of iodine molecules
164 Dye-Sensitized Solar Cells

adsorbed at the Cul surface, acting as hole-trapping sites [6.3]. DSCs incorporating
CuSCN as the hole-transporter have demonstrated improved stabilities; CuSCN does
not decompose to SCN- and there is no indication of excessive SCN- creating surface
traps. However, these DSCs exhibit slightly lower efficiencies of ~2 % under 1 sun
illumination [6.4-6.6]. Crystallization of the copper compounds in the pores is also
thought to contribute to the cell degradation. Another class of materials constituting
effective hole conductors are organic molecular and polymer semiconductors. Though
the charge mobilities in organic hole-transporters are generally lower than in cop­
per-based compounds, the mobilities are still suitably high. Moreover, the molecules
are solution processable, rendering them well suited for mesopore infiltration, they
should be cheap, and innumerable variations can be conceived and created. For all
these reasons, organic semiconductors are highly versatile. The first solid-state DSCs
incorporating polymeric and molecular organic hole transporters were realised in 1997
and 1998 respectively [6.7, 6.9]. The best performing solid-state DSCs incorporating
organic hole transport materials deliver over 5 % solar-to-electrical power conversion
efficiency [6.8].
Here, we review the development of solid-state dye-sensitized solar cells incor­
porating molecular hole transporters. We discuss the key steps which have been taken
in the development thus far, describing the details of the photovoltaic process in a gen­
eral and precise form. The advantages of this system are highlighted, and we further
discuss the short falls and required future progress necessary to enable this concept to
lead the next generation of photovoltaics.
The key components to the solid-state DSC are the sensitizer, electron-
transporting mesoscopic metal oxide electrode, and the molecular hole-transporter.
Similarly to the liquid electrolyte-based DSC, the sensitizers are typically ruthenium
bypyridyl complexes, and are adsorbed "uniformly" upon the surface of the mes­
oscopic electrode. The mesoscopic electrode is generally mesoporous Ti02, fabri­
cated from sol-gel-processed and sintered nanoparticles. Currently, the most effective
hole-transporter is a methoxy triaryl diamine substituted spiro centered molecule,
namely 2,2/,7,7/-tetrakis(A^,A^-di-/?-methoxyphenylamine)-9,9/-spirobifluorene (spiro-
OMeTAD), which is infiltrated into the mesoporous Ti0 2 via solution casting [6.9].
Light is absorbed in the sensitizer, promoting an electron from the highest occupied
molecular orbitals (HOMOs) to the lowest unoccupied molecular orbitals (LUMOs)
with subsequent electron transfer to and migration in the Ti0 2 and hole-transfer from
the HOMO level of the dye, to the HOMO of the hole-transporter and migration in the
hole-transporter. The whole device is mounted upon glass coated with a transparent
conducting oxide, typically fluorine doped tin oxide (FTO), which carries the electron
away to the external circuit. A flat compact Ti0 2 interlayer is inserted between the
FTO and the mesoporous Ti0 2 to block direct contact of the FTO electrode with the
hole-transporter. Contact to the hole-transporter is made with a high work function
metallic electrode, typically gold or silver. A schematic of the device structure, along
with a representative energy level diagram is presented in Figure 6.1.
When first realized by Bach and Grätzel in 1998, the molecular hole-transporter
based solid-state DSC had an efficiency of less than one percent [6.9]. But with a
concerted effort of only a few research groups, the efficiency has been pushed up
to 5 % in the current best-performing systems [6.8]. Figure 6.2 shows the efficiency
milestones as a function of time.
Solid-state dye-sensitized solar cells 165

In the following sections, we go through the various components of the solid-state


DSC in detail, describing how they influence the solar cell operation, how each compo­
nent has been modified to improve solar cell performance and how we have furthered our
understanding of the photovoltaic process. The unresolved issues are also highlighted.

6.2 SPIRO-OMeTAD-BASED SOLID-STATE DYE-SENSITIZED


SOLAR CELL

In 1998, Bach and Grätzel made a miraculous development and presented a DSC with
0.7 % power conversion efficiency incorporating the solid organic molecule

Fig. 6.1 (a) A schematic illustration of a solid-state DSC. (b) An energy level diagram repre­
senting the Ti02-dye-hole-transporter heterojimction (energy scale w.r.t. vacuum) indicating the
main processes of excitation, electron injection (or transfer) hole injection (or dye regeneration)
and recombination between the electrons in the Ti0 2 with the holes in the hole-transporter.
166 Dye-Sensitized Solar Cells

Fig. 6.2 The power conversion efficiency measured under AM 1.5 simulated sun light,
(1) Original concept introduced by Bach et al., intensity 9.4 mW cm-2. (2) Optimization of
lithium salt and tBP content by Kruger et al., intensity 100 mW cm-2, verified at NREL. (3)
Dye adsorption in the presence of silver with subsequent iodide rinsing by Kruger et al., inten­
sity 100 mW cm-2. (4) Use of an amphiphilic ruthenium complex by Schmidt-Mende et al.,
intensity 100 mW cm-2. (5) Use of a high extinction coefficiecnt indolene-based sensitizer by
Schmidt-Mende et al., intensity 100 mW cm-2. (6) Use of an ion-coordinating sensitizer com­
bined with reflective silver electrodes by Snaith et al., intensity 126 mW cm-2.

2,2/,7,7/-tetrakis (A^,A/-di-/?-methoxyphenylamine)-9,9/-spirobifluorene (spiro-OMeTAD)


as the redox couple and hole-transport component [6.9]. A decade on and spiro-OMeTAD
remains the hole-transporter of choice. The reasons for its suitability will become clear
throughout this chapter, though we do not believe it to be ideal. The chemical structure
and original current-voltage (JV) curves from the initial work of Bach et al. [6.9] and a
cyclic voltammogram of spiro-OMeTAD are presented in Figure 6.3.

6.3 THE INFLUENCE OF ADDITIVES UPON THE SOLAR


CELL PERFORMANCE

In order to obtain the highest efficiency for the solid-state DSC, two specific additives
to the hole-transporter phase are required: 4-^ri-butylpyridine (tBP) and an ionic
salt, typically lithium-bis(trifluoromethanesulfonyl)imide ( (Li-TFSI) [6.11, 6.12]. In
Figure 6.4, the current voltage characteristics for solid-state DSCs incorporating neat
spiro-OMeTAD, with the addition of Li-TFSI on its own and tBP + Li-TFSI are shown,
measured under simulated AM 1.5 sun light of 100 mW cm - 2 .
It is clearly apparent that, without the additives, the device is exceptionally
unremarkable. The precise function of the additives will be discussed in detail in
the respective sections below. In brief, tBP helps to solvate the lithium salt and causes
Solid-state dye-sensitized solar cells 167

Fig. 6.3 (a) Current-density/voltage characteristics for the first efficiently operating solid-
state DSC incorporating a molecular hole-transporter presented in reference [6.9], obtained
in the dark (I) and under white-light illumination at 9.4 mW cm -2 (II). The spectral distribu­
tion corresponded to global air mass 1.5 corrected for spectral mismatch. The short-circuit
current density was 0.32 mA cm-2, the open-circuit voltage 342 mV and the fill factor 0.62,
corresponding to an overall conversion efficiency of 0.74 %. (b) A cyclic voltammogram of
spiro-OMeTAD in CH2C12/0.1 M [N(n-C4H9)4](PF6) in dichloromethane versus [FeII(Cp)2/
FeIII(Cp)2] at room temperature; scan rate 100 mV/s. The redox potential of ferrocene
E[FeII(Cp)2/FeIII(Cp)2] in organic solvents is typically 700 m V vs. NHE. Two one-electron
oxidations and a two electron oxidation can be observed at 118 (le~), 232 (le~) and 436 m V
(2e~) vs. ferrocene respectively (equivalent to 818, 932 and 1136 mV vs. NHE and approxi­
mately 5.3, 5.4 and 5.6 eV vs. vacuum). Reproduced with permission from reference [6.10].
The chemical structure of spiro-OMeTAD is also shown.

a positive shift in the T i 0 2 conduction band with respect to the hole-transporter energy
levels. The positive shift is caused by a weak binding of the pyridine to the T i 0 2
surface, which donates a negative charge to the surface, and by acting as a "pro­
ton sponge" removes protons from the titania surface, with the protons themselves
potential-determining. Furthermore, the fBP also interacts with the Li-TFSI, inhibiting
it from absorbing to the surface, with the Li salt also being potential determining. The
original device of Bach et al. contained no fBP and a fully protonated sensitizer (4 pro­
tons), which is why it exhibited a low open-circuit voltage (342 mV). The function of
168 Dye-Sensitized Solar Cells

Fig. 6.4 Current-voltage curves for a solid-state DSC using "Z907" as the sensitizer, incor­
porating pure spiro-OMeTAD, with the addition of Li-TFSI on its own and with the addi­
tion of both Li-TFSI and iBP, measured in the dark and under AM 1.5 simulated sun light of
100 mW cm-2.

the ionic salts is complex and not entirely understood: The addition inhibits charge
recombination, stabilizing the photogenerated charge in the device. Furthermore, it
increases the conductivity in the hole-transporter phase, appears to increase the charge
carrier mobility and possibly enhances the charge generation efficiency.

6.4 CHARGE GENERATION: ELECTRON TRANSFER

The photo-induced electron transfer process from excited sensitizer molecules into
Ti0 2 is still a topic of fierce debate, despite it having been studied for two decades.
One of the main issues in this field is that there is no standard Ti0 2 with which to make
comparisons. Ti0 2 , or more accurately Ti02-x, is horrendously non-stoichiometric,
and the stoichiometry depends strongly upon the material fabrication and post treat­
ments. Moreover, it significantly influences the electronic nature of the material. The
level of carbon contamination will also vary from batch to batch and is likely to be
influential. Furthermore, the system is so complex that changes in the purity of the
dye, film dying and subsequent device fabrication protocols may significantly influ­
ence the electronic processes.
For dyed Ti0 2 electrodes, in-filled with an "inert" electrolyte, the electron
transfer from an excited Ru-complex to the Ti0 2 is usually observed to be biphasic,
with an "ultra fast" component (within a few tens of fs) accounting for between 35 to
95 % of the signal and a slower component (1 to 100 ps, sometimes ns) accounting
Solid-state dye-sensitized solar cells 169

for the rest [6.13]. It has been proposed that the slow phase arises from the dye being
absorbed to energetically inhomogeneous sites on the Ti0 2 surface [6.14]. However,
Wenger et al. very convincingly demonstrated that, for dye adsorbed sparsely to the
Ti0 2 surface, the electron transfer was almost entirely ultrafast [6.13]. The slower
transfer rate appeared to only occur when dye-dye interaction or aggregation took
place. The ultrafast transient adsorption traces, probing the electron injection kinetics
for mesoporous Ti0 2 sensitized with "Ν3" from high and low concentration solutions
are presented in Figure 6.5.
We note that in order to obtain the monophasic ultrafast photoinduced elec­
tron transfer from a commercially available N3 sensitizer, as the one presented in
Figure 6.5, the dye loading has to be reduced to such an extent that the optical density
of a 6-μιη thick film is only 0.34 at peak wavelength [6.13]. This is below practical
value and suggests that real commercial devices may have to be engineered to accom­
modate the slow phase electron transfer. Furthermore, the electrolyte system and the
solid-state hole-transporter are likely to interact with the dye and may very well influ­
ence the electron-transfer kinetics and the fraction of dye undergoing slow as opposed
to ultrafast electron transfer.
The electrolyte contains many components which can tune the surface potential
of the Ti0 2 , and in principle reduce or increase the speed and efficiency of electron
transfer, with the ability to maximize final solar cell efficiency. Incorporating compo­
nents such as lithium salts to the electrolyte tends to cause a positive shift of the Ti0 2
conduction band with respect to the electrolyte potential, increasing current genera­
tion efficiency but reducing cell voltage. The addition of components such as pyridine
results in a negative shift of the conduction band with respect to the electrolyte poten­
tial, reducing the current generation efficiency but increasing the cell voltage. Durrant
and co workers have shown that the electron injection can be rather slow in certain
instances, even in the ns regime [6.15, 6.16]. This then becomes comparable with the
excited state lifetime (a few to tens of ns) resulting in a non-negligible loss in electron

Fig. 6.5 Transient absorbance signals measured at 860 nm upon excitation at 535 nm of the
commercial N3 dye adsorbed on Ti0 2 films from 3 x 10~4 M (O) and 1.5 x 10~5 M (Δ) solutions
in ethanol. Drawn with data from reference [6.13].
170 Dye-Sensitized Solar Cells

transfer efficiency. Even in "optimized" electrolyte-based cells, it is apparent that the


current generation efficiency is sensitive to the Li salt content in the electrolyte [6.17].
The sensitivity of charge generation and photocurrent to electrolyte composition is
depicted most elegantly in a recent study by Koops et al., where they indeed demon­
strated such a tuning of the electron injection rate, charge generation and open-circuit
voltage by varying the Li-TFSI and fBP content in the electrolyte [6.18].
This also holds true to a certain extent for the solid-state cell. However, tuning
the charge generation and open-circuit voltage is not as straightforward. Both fBP and
Li-TFSI have been added to the hole-transporter matrix at an optimized content, and
as can been seen in Figure 6.4, they indeed alter the current and voltage as expected.
However, they exert a more convoluted influence upon many processes in the solid-
state DSC and are thus not the ideal components to vary in order to optimize electron
transfer. fBP certainly augmented the cell open-circuit voltage, by up to 200 mV in
comparison to a cell with exclusively Li-TFSI does indeed cause a reduction in the
cell open-circuit voltage when added on its own. However, in combination with fBP,
adding Li salts to the solid-state matrix usually results in an increase in cell open-cir­
cuit voltage, contrary to its influence in the liquid cell. Hence, in the complete mix,
the tricks played with the electrolyte system are not directly mapped to the solid-
state cell. In Figure 6.6, we present the open-circuit voltage and short-circuit current
versus Li-TFSI content in the hole-transporter matrix. Upon the addition of a small
amount of Li-TFSI the short-circuit current increases considerably, however upon
further increase, it remains more or less constant, indicating a limited ability for Li-
TFSI to augment the current generation efficiency. In contrast, the open-circuit volt­
age was monotonically raised with an increasing lithium salt content, indicating that
the lithium does not cause a positive shift of the T i 0 2 conduction band. In Figure 6.6,

Fig. 6.6 The short-circuit current and open-circuit voltage plotted as functions of the Li-TFSI
concentration (in comparison to spiro-OMeTAD) in a solid-state DSC using a K51 sensitizer
and the standard concentration of iBP throughout. Please not that the points on the Y axis at
1 % concentration correspond to no Li-TFSI. Adapted from reference [6.17].
Solid-state dye-sensitized solar cells 171

the device presented included a sensitizer containing ion-coordinating side chains


(K51) [6.19]. This exaggerates the increase in Voc with lithium salt concentration for
reasons that will be addressed later, though the augmentation of the open-circuit volt­
age with Li salt concentration is generally observed with most sensitizers employed
in solid-state DSCs [6.11, 6.17]. An area where further improvements are required
is in devising routes to fine tune the Ti0 2 energy level with respect to the dye and
hole-transporter, such as using coadsorbed dipolar molecules [6.20, 6.21], since the
additive route presently employed is rather clumsy.

6.5 REDUCTIVE QUENCHING

There have been relatively few studies of the charge generation in solid-state hole-
transporter-based DSCs. The primary reason is that there is only a small number
of researchers working in this area. A recent study of an indolene-based organic
sensitizer, termed D149, employed in solid-state DSCs, appears to demonstrate
that reductive quenching can constitute the main charge generation mechanism
for certain dye-hole transporter combinations [6.22]. Figure 6.7 portrays the time-
resolved photoluminescence decay for dye-sensitized Ti0 2 and dye-sensitized Ti0 2
infiltrated with spiro-OMeTAD (with and without Li-TFSI). The fitting of these
decays reveals that the electron transfer to the Ti0 2 took place in approximately
300 ps whereas hole-transfer occurred in only 40 ps, implying that over 90 % of
the excited dye was reductively quenched [6.22]. Interestingly, the addition of Li-
TFSI to the hole-transporter did not influence the photoluminescence decay life­
time, which suggests that it did not enhance the electron transfer rate or efficiency.
Alternatively, that the electron-transfer was still much slower than the hole-trans­
fer, with the latter being uninfluenced by the presence of the Li-TFSI. The proc­
esses and relative rates are also illustrated in Figure 6.7. Many more studies need
to be performed to understand whether this reductive quenching mechanism is
general for solid-state DSCs, and if it is a major loss mechanism. However, if it
occurs with many sensitizers, we need to consider how the electron will transfer
from the reduced dye to the Ti0 2 and how this competes with recombination with
the oxidized hole-transporter. A study of DSCs based on ionic liquids revealed that
a surprisingly fast reductive quenching process was turned on with pure iodide
melts, however this channel was observed to be an unproductive route for photo-
current generation [6.23].

6.6 CHARGE GENERATION: HOLE-TRANSFER

Though we have observed reductive quenching for a specific system, it is generally


considered that, for efficiently operating DSCs, electron transfer from the photoex-
cited dye into the Ti0 2 conduction band is the initial step in the charge generation
process. Following electron transfer from the excited dye into the Ti0 2 , the oxidized
dye must be regenerated via hole-transfer from the HOMO of the dye to the HOMO
172 Dye-Sensitized Solar Cells

Fig. 6.7 (a) The time-resolved photoluminescence decay for mesoporous Ti0 2 sensitized with
D149, and that infiltrated with pure spiro-OMeTAD and spiro-OMeTAD containing Li-TFSI
and tbp. (b) A schematic energy level diagram for the Ti02/dye/spiro-OMeTAD heterojunction,
with the lifetimes for the relative electronic processes marked. r nr is the lifetime for the non-
radiative decay channel in the dye whereas rrad is the lifetime for the radiative decay channel.
Drawn and adapted with data from reference [6.22].

of the hole-transporter. This "interception" reaction must occur prior to recombina­


tion of the electron in the T i 0 2 with the hole on the sensitizer (back reaction). The
"back reaction" typically takes hundreds of microseconds to several milliseconds,
leaving plenty of time for dye-regeneration. Bach et al. studied the dye regenera­
tion kinetics in solid-state DSC composites using a bipyridyl ruthenium complex,
cw-RuL2(SCN)2 (where L is 4,4 / -dicarboxy-2,2 / -bipyridyl) (termed N719) as the
sensitizer and spiro-OMeTAD as the hole-transporter, via transient photoinduced
absorption spectroscopy. They observed sub-nanosecond dye regeneration, indicative
Solid-state dye-sensitized solar cells 173

of the process being very efficient. Figure 6.8 depicts the transient absorption trace
at 520 nm, probing the transition from the bleached dye signal to the oxidized spiro-
OMeTAD signal (i.e., dye regeneration).
This "one-electron" dye regeneration mechanism is quite dissimilar to the proc­
ess occurring in the iodide/triiodide-based electrolyte system, where four iodide mol­
ecules and two holes are involved in each regeneration reaction. As a consequence, a
lower driving force for the reaction should be necessary for the hole-transporter-based
system, which could enable a closer matching of the dye and hole transporter HOMO
energy levels, thus ensuing higher open-circuit voltages [6.24]. In fact, with spiro-
OMeTAD, the open-circuit voltages of the devices are usually very high in compari­
son to liquid iodide/triiodide-based cells, and open-circuit voltages well in excess of
1 volt have been observed [6.25]. This is primarily due to the HOMO level of spiro-
OMeTAD being significantly more positive than the redox potential of iodode/trii-
odide. Garcia-Canadas et al. have estimated the relative offset in the energy levels at
the heterojunction by impedance and cyclic voltammetry measurements [6.26]. They
found that when using N719 (Ru(4,4 / -dicarboxy-2,2 / -bipyridyI) 2 (SCN) 2 ) as the sensi-
tizer, there was only a 250 mV offset between the center of the spiro-OMeTAD HOMO
level and the dye HOMO level, indicating near optimum matching for dye regenera­
tion and maximizing open-circuit voltage. This is illustrated in Figure 6.8. We do note
however, that based on the sub-ns dye regeneration for this system, it is likely that the
HOMO of the hole-transporter can be pushed closer to that of the dye without losses to
the dye regeneration efficiency.

Fig. 6.8 (a) Transient absorption kinetics obtained for dye-sensitized Ti0 2 films covered with
spior-OMeTAD (o, top trace) and with PC/EC (Δ, bottom trace) at a probe wavelength of
520 nm after excitation at 602 nm with pulse intensities of 60 μΐ/cm2. The solid lines show
results of exponential fitting to the data. For the sensitized Ti02/spiro-OMeTAD heterojunction,
a double-exponential fit yielded component lifetimes of 100 and 2300 ps [6.86]. Reprinted with
permission from /. Am. Chem. Soc, 32, 7445 (1999). Copyright 2005 American Chemical
Society, (b) A schematic of the energy levels of the main components of the Ti02/Dye/spiro-
OMeTAD system as derived by Garcia-Canadas et al., redrawn from reference [6.26].
174 Dye-Sensitized Solar Cells

6.7 CHARGE TRANSPORT IN MOLECULAR


HOLE-TRANSPORTERS

In contrast to inorganic semiconductors, the charge transport in molecular solids and


semiconducting polymers cannot generally be considered as band-like. There are vari­
ous models to describe the mechanism of charge transport in organic semiconductors.
The three main mechanisms are polaron formation and transport, carrier trapping by
impurities or defects, and energetic disorder leading to an energetically broad distri­
bution of density-of-states, typically Gaussian [6.27, 6.28].
The presence of an extra electron or the absence of an electron (presence of a
hole) on an organic semiconductor results in a molecular relaxation due to the strong
electron-phonon coupling in these materials. This induces "self-trapping" in a carrier
and creates a quasi-particle known as a polaron. In the simplest terms, a polaron is a
charge coupled with a lattice distortion, and it is this entire quasi-particle that must
be displaced in order for charge transport to occur. In an organic semiconductor, the
extent of the wave function associated with the polaron is comparable to the separa­
tion of molecules, and is thus termed a 'small polaron'. It has an intrinsic binding
energy which must be overcome for the carrier to move to an adjacent site, with a
characteristic activation energy of typically half the binding energy [6.29]. This acti­
vation energy has been calculated to be 0.145 eV for poly [methyl (phenyl)silylene]
[6.30], giving a polaron binding energy of 0.29 eV. However, this binding energy is
expected to vary considerably between materials, largely dependent upon the spatial
extent of the polaron and the degree of lattice distortion.
An alternative mode of transport is trap-controlled transport, or the multiple-
trapping model as used to describe the charge transport in mesoporous Ti0 2 . This
considers all charges in sites below certain energy to be trapped, and all those in sites
above to be free. De-trapping occurs through a thermally assisted process with the
free charge moving through the lattice until it falls into the next trap site. Both polaron
and trap-controlled transport tend to have a temperature dependence of mobility (μ)
of In μα 1/T. It is thus difficult to experimentally distinguish between the two [6.29].
The hopping model, for a polaron-free system, considers an energetic distribu­
tion of sites, with the hopping probability proportional to an electronic wave function
overlap factor and a Boltzmann factor for jumps to higher energy sites. The original
version of the hopping model is the Miller-Abrahams jump rate (1960) [6.31]. This
was extended by Mott to include variable range hopping (VRH) where carriers can
hop over a small distance with a large activation energy or a large distance to a site
with similar or lower energy [6.32]. The temperature dependence of this VRH mobil­
ity lies between In μ α 1/T 2 and 1/T 4.
Even though the temperature dependence for the hopping across a Gaussian
density-of-states (energetic disorder) is significantly different from that of the polaron
transport and trap-controlled transport, it remains difficult to experimentally dis­
tinguish between the three due to the relatively small temperature range accessible
experimentally [6.29]. There has been much work focusing on refining the transport
theories, and comparing calculated models with experimental data. Fishchuk et al.
have shown very good agreement between data and theory by superimposing polaron
Solid-state dye-sensitized solar cells 175

and disordered effects [6.30]. For a real organic semiconductor, the charge transport
is likely to be a superposition of the three effects (polaron, disorder and trapping) with
disorder being more significant at low temperatures, polaron effects at high tempera­
tures and trapping occurring when there is a high density of defects or impurities.

6.8 HOLE MOBILITY IN SPIRO-OMeTAD

In the solid-state DSC, the role of the hole-transporter is to receive holes from the
oxidized dye and transport them out of the device. The hole-mobility in the material
is thus a critical parameter. For an amorphous organic semicondutor, spiro-OMeTAD
has a moderately high hole mobility of ~10~4 cm2/Vs as measured by the time-of-
flight technique [6.33]. A time-of-flight mobility device is composed of a thick semi­
conductor film (1 to 10 μιη) sandwiched between two electrodes. In a time-of-flight
measurement, a packet of charge is photo-generated at one side of the film and sub­
sequently drifts across the film under an applied bias. While it drifts, a current flows
in the external circuit, and this current starts to decay once the first carriers reach the
far electrode, decaying to zero after all carriers have recombined. If the charge trans­
port is non-dispersive, then a flat plateau is observed while all the carriers uniformly
traverse the film, and the current quickly decays to zero once they reach the far side. If
the charge transport is highly dispersive, then no plateau is seen and a constant decay
in the transient current trace is observed, rendering it difficult to use this technique to
characterize semiconductors with highly dispersive charge transport. In Figure 6.9,
a typical time-of-flight hole transient for spiro-OMeTAD is presented [6.33]. The
clearly visible "plateau" followed by a "knee" is a characteristic of non-dispersive
hole transport, implying that the energetic disorder is relatively low and that there is
not a large density of defect trap sites within the material.

6.9 INFLUENCE OF CHARGE DENSITY ON THE


HOLE-MOBILITY IN MOLECULAR SEMICONDUCTORS

Vastly dissimilar charge carrier mobilities can be estimated for the same organic semi­
conductors depending upon which techniques are used to make the measurements.
For instance, mobilities derived from space-charge-limited current measurements are
usually many orders of magnitude lower than those derived from measurements in
field effect transistor configuration [6.34]. The physical meaning of mobility, i.e., the
ratio of the drift velocity of an ensemble of carriers to the electric field, should be the
same for all methods. The main difference in mobility derived from various experi­
ment techniques has been shown to be due to the greatly different charge densities in
the studied materials for each technique, and a mobility which often depends strongly
upon the charge density [6.35]. The charge density dependence arises from energetic
disorder and with a broad distribution of states. If we consider a globally Gaussian
energetic distribution of states, then near the center of the density of states (DoS), the
states are close enough to enable facile inter-site hopping and the mobility for charges
176 Dye-Sensitized Solar Cells

Fig.6.9 Atypicalroom-temperatureTOFholetransientinspiro-OMeTAD((i = 0.95 μιη,ν57 V).


The transit time ttr = d/μΕ of around 10 ^s gives a mobility of ~ 1 x 10~4 cm2/Vs. The inset
shows a schematic of a diode used for time-of-flight measurements. Data for the main figure
reproduced with permission from reference [6.33].

in this region is high. Near the tail of the DoS distribution, the inter-site distance is
large and hopping is relatively restricted. As the charge density increases, the lower
energy sites become predominantly filled and the average occupied site energy moves
closer to the facile transport region, resulting in an increase in the average mobility.
This is schematically illustrated in Figure 6.10. The calculated hole mobility in two
semiconducting polymers is also shown as a function of charge density, for the low
charge density, diode configuration and the high charge density field-effect transis­
tor configuration. When it comes to measuring mobility, the time-of-flight technique
is most sensitive to the most mobile carriers and hence mobilities similar to those
derived from field effect measurements are usually observed.
The charge density in a space-charge-limited diode is typically on the order
of 1015 to 1016 cm -3 as compared to field effect transistors where charge densities in
excess of 1018 cm -3 are generally estimated [6.34]. For a dye-sensitized solar cell, the
charge densities are in the region of 1016 to 1018 cm -3 depending upon light intensity
and operating conditions, typically a few times 1017 cm -3 under full sun illumination.
For this reason, the DSC is in an intermediate charge density regime, but further to
this, the influence of the meso-confinement, imposed by the Ti0 2 , upon charge mobil­
ity is unclear. It is therefore a challenge to estimate the charge carrier mobility in a
system and under conditions closely matching those in the solid-state DSC. Snaith
et al. have estimated the mobility through molecular hole-transporters infiltrated into
dye-sensitized mesoporous Ti0 2 , by measuring the conductivity (σ) through in-plane
transistor-type devices under illumination, estimating the charge density (n) and using
the simple relation between conductivity, charge density and mobility (μ), σ = μηβ,
to derive the latter [6.36, 6.37]. Figure 6.11 presents the hole-mobility versus the inci­
dent light intensity for a triphenylamine trimer and spiro-OMeTAD infiltrated into
Solid-state dye-sensitized solar cells 177

Fig. 6.10 (a) A schematic representation of the density of states distribution in an organic semi­
conductor, the energy region in which facile charge transport can occur is identified as Et. The
dashes represent available sites, and the open-circles represent sites occupied with holes. The
energy scale is representative of a hypothetical hole-transporter, (b) The mobility as a function
of the hole density, p, in a diode andfield-effecttransistor for P3HT and OCiCi0-PPV. Part (b)
is reproduced with permission from ref. [6.34], Copyright 2003 American Physical Society.

dyed mesoporous Ti0 2 . For the triphenylamine trimer infiltrated into mesoporous
Ti0 2 , a strong dependence of the hole mobility to incident light intensity and hence
charge density was observed. However, for spiro-OMeTAD, the hole mobility was
found to be less sensitive to changes in light intensity and charge density [6.36, 6.37].
This therefore suggests that for the currently best-performing system, variations in
hole mobility with charge density are only minor. However, this cannot be generalized
to all hole-transporters, and for future successors of spiro-OMeTAD, it may need to
be considered.

6.10 THE INFLUENCE OF CHEMICAL P-DOPING UPON


CONDUCTIVITY AND HOLE-MOBILITY

Further to the confinement imposed by the mesoporous Ti0 2 discussed above, the
hole-transporter operates most efficiently when it contains a certain fraction of ionic
additives and coordinating solvents, see Figure 6.4. Until recently, the usual cocktail
has also included a chemical p-dopant, typical an antimony salt [6.9, 6.12]. Originally,
it was considered necessary to partially oxidize the hole transporter to allow injection
of electrons from the cathode into the vacant "doped" holes in the HOMO level of
the hole-transporter. However, if we consider the independent flow of electrons and
holes, then once the electron has been injected into the Ti0 2 , the oxidized dye can be
regenerated directly via hole transfer to the neutral hole-transporter HOMO levels.
The hole can then be transported through the hole-transporter and collected at the
metal cathode without any requirement for prior oxidization.
178 Dye-Sensitized Solar Cells

Fig. 6.11 (a) The hole-mobility versus the incident light intensity for a triphenylamine trimer
(chemical structure shown as inset) and spiro-OMeTAD infiltrated into dye-sensitized mesopo-
rous Ti02 measured in an in-plane configuration, (b) A schematic illustration of the in-plane
conductivity device. Figures adapted from references [6.36, 6.37].

Aside from whether it is fundamentally necessary to dope or not, two things


could conceptually be improved via the addition of chemical p-dopants. First, the
mobility may be increased due to the augmented charge density, given a charge-den­
sity-dependant mobility, as discussed above. Secondly, it should be possible to reduce
the series resistance through the cell, and hence improve the fill factor by simply
increasing the conductivity. This is specifically important when there is a thick "cap­
ping" layer of hole-transporter on top of the mesoporous film [6.38]. Detrimental
influences include: (i) increasing the recombination rate between electrons in the Ti0 2
with holes in the hole-transporter, simply due to the increased hole density [6.12,
6.39], (ii) the inclusion of the chemical oxidant influencing the film-forming prop­
erties and morphology of the hole-transporter, thus affecting the pore filling, (iii)
the introduction of dopant ions and counter ions increasing the energetic disorder
in the material to such an extent that the mobility is actually reduced [6.40]. For
spiro-OMeTAD these negative influences appear to out-weigh the possible advan­
tages, making it disadvantageous to p-dope the material in the current state-of-the-art
devices. In Figure 6.12, the photovoltaic performance parameters for a standard solid-
state DSC incorporating a range of antimony salt chemical p-dopant concentrations,
are shown. The most efficient devices contained no oxidant, although the performance
parameters only varied marginally. The same trends have been observed with various
dopants [6.12]. Even though spiro-OMeTAD has not benefited from chemical oxidi­
zation, for other hole-transporters, especially those exhibiting a strong dependence of
the mobility upon charge density, i.e., highly dispersive charge transport characteris­
tics with a large degree of energetic disorder, it may be advantageous to dope at an
optimized concentration.
When considering the concentration with which to chemically p-dope the mate­
rial, the previously standard composition included only approximately 0.16 % dopant
to spiro-OMeTAD molecules. It is surprising that such low concentrations have any
Solid-state dye-sensitized solar cells 179

Fig. 6.12 Device performance parameters for solar cells tested under AM 1.5 simulated solar
illumination at 100 mW cm-2, with a range of concentrations of antimony (Sb) salt in the spiro-
OMeTAD casting solution. The % concentration is with respect to spiro-OMeTAD molecules, not
that the points on the y-axis correspond to devices without dopant. Figure drawn with data from
ref. [6.12].

positive influence upon conductivity or mobility. Jiang et al. observed a very strong
dependence of the charge carrier mobility upon dopant concentration when poly(3-
hexathiophene) was electrochemically doped [6.41]. At concentrations below 1 %,
there was actually a reduction in the mobility, leading to a sudden and very rapid rise
upon further increase in dopant concentration. This data is presented in Figure 6.13,
and the phenomenon can be explained by understanding that although chemical dop­
ing of an organic semiconductor does increase the charge density, it also introduces
the same number of deep Coulomb traps associated with the counter ions. At low
doping levels, the Coulomb traps tend to reduce the charge carrier mobility, whereas
at increased levels, the potential wells surrounding the counterions begin to overlap
and the mobility can increase considerably [6.42, 6.43]. The influence of the coulomb
traps upon the potential landscape at low and high concentrations is shown schemati­
cally in Figure 6.13.
Arkhipov et al. demonstrated theoretically that mobility enhancements can
only be expected upon chemical doping of a material if it is strongly disordered (for a
Gaussian distribution, σ > 0.15 eV) and/or under strong electric fields. If the system is
less disordered, carrier localization in the increasing density of Coulomb traps domi­
nates and the mobility drops with an increasing doping [6.40]. The energetic disorder
parameter (deviation of the Gaussian DoS) for the HOMO level of spiro-OMeTAD
has been measured to be only 65 meV [6.26]. Accompanied by the non-dispersive
transport observed by Poplavskyy et al. [6.33], this implies that an enhanced mobility
through chemical oxidization is unlikely to ever be achieved with spiro-OMeTAD,
though an increased conductivity may be obtained. For other hole-transporters, the
story may be very different though. The non-generality of this is illustrated by recent
work where a room temperature molten molecular hole transporter was used in a dye-
sensitized solar cell, and it required over 10 % chemical p-doping to have sufficient
conductivity and operate effectively in the DSC [6.44].
180 Dye-Sensitized Solar Cells

Fig. 6.13 (a) The mobility versus the doping level (with respect to the P3HT monomer con­
centration) for P3HT electrochemically oxidized. The legend values correspond to % regio
regularity [6.41]. (b) and (c) The influence of coulomb traps introduced by dopant counterions
upon the poltential landscape at low (b) and high (c) doping densities [6.42].

6.11 THE INFLUENCE OF IONIC SALTS ON CONDUCTIVITY


AND HOLE-MOBILITY

Lithium salts appear to have a phenomenal influence upon the processes occur­
ring in the solar cell, not least of all on the charge transport and conductivity in the
hole-transporter. Table 6.14 lists the conductivity of spiro-MeOTAD infiltrated into
mesoporous Ti0 2 with a range of additives present. The conductivity of neat spiro-
MeOTAD was relatively low, 2.5 x 10~7 S cm-1, and remained so even with the addi­
tion of 0.16 % doping by an antimony salt. This is not surprising considering the
discussion in Section 6.10. Upon the addition of 12 % lithium TFSI doping, the con­
ductivity is enhanced by close to two orders of magnitude, and a small further incre­
ment is observed with subsequent addition of chemical oxidant.
Solid-state dye-sensitized solar cells 181

Table 6.14 The conductivity of spiro-OMeTAD infiltrated into mesoporous Ti0 2 : neat, with
the addition of antimony salt chemical p-dopant, with the addition of Li-TFSI and with both
additives. iBP was always present with Li-TFSI but has a negligible influence on its own. Taken
from reference [6.12].

Device Conductivity (S cm-1)

Neat spiro-OMeTAD 2.5 x 10~7


Spiro-OMeTAD + antimony salt 2.2 x 10~7
Spiro-OMeTAD + Li-TFSI 2.0 x 10~5
Spiro-OMeTAD + antimony + Li-TFSI 3.0 x 10~5

The enhanced conductivity with the addition of Li-TSFI appeared not to be ionic
since it was persistent. The enhancement could be due either to an increased charge den­
sity and/or an enhanced mobility. Interestingly, the Li-TFSI did not chemically oxidize
the spiro-OMeTAD, which became apparent when examining the UV-Vis absorption
spectra of thin films or in solution, shown in Figure 6.15. We can therefore rule out
chemical oxidization as a means by which the charge density is increased. However,
in a device configuration where there are metallic contacts, the TFSI counterions may
stabilize a large density of positive space-charge under current flow, which could con­
tribute towards the increased conductivity. There does however appear to be close to one
order of magnitude increase in the hole-mobility with the addition of the lithium salts,
as estimated by the time-of-flight technique, in space-charge-limited diodes and in in-
plane conductivity devices [6.12, 6.37]. It is not intuitive that the enhancement in mobil­
ity should occur with the addition of lithium salts. The addition of 12 % ionic species
should considerably alter the potential landscape similarly to the effect of electrochemi­
cal doping. The potential profile between two ions separated by a distance b is,

(6.1)

If the ions are uniformly distributed at 12 % concentration, the inter-ion distance


(b) is only 2.4 nm. Considering a dielectric constant (ε) of 3 for spiro-OMeTAD, the
Coulomb traps associated with each counterion will overlap considerably. Figure 6.15
displays the calculated potential profile between two counterions in spiro-OMeTAD
with 12 % ionic doping. The "potential barrier" for detrapping of a charge from the cou­
lomb well has become reduced by 0.8 eV at this concentration, as compared to charge
associated with isolated counterions. It is quite apparent that the intrinsic disorder in
the hole-transporter would be entirely dominated by the extrinsic disorder induced
by the addition of the salts. Even though this should not necessarily lead to enhanced
mobility, it will clearly have a significantly influence on the charge transport.

6.12 CURRENT COLLECTION

Once an electron and a hole are separated across the Ti0 2 -dye-spiro-OMeTAD het-
erojunction, they must diffuse apart and be collected at their respective electrodes
182 Dye-Sensitized Solar Cells

Fig. 6.15 (a) The chemical structure of Li-TFSI and antimony salt chemical p-dopant. (b) UV-
Vis absorption spectra of a solution of neat spiro-OMeTAD, with the addition of 12 % Li-TFSI
and with 0.16 % antimony salt chemical oxidant. (c) A calculated potential profile of two point
charges 2.4 nm apart in a material with a dielectric constant of 3. The solid and dashed lines
represent the potential profile of isolated ions and the dotted grey line is the net potential profile.
Adapted from reference [6.12].

before recombining with either each other or any other charge. The figure of merit by
which we can gain some quantitative appreciation of the charge collection efficiency
is the diffusion length, LD, which is simply the square root of the product of the charge
diffusion coefficient (De) and lifetime (r e ), (LD = yJDje ). If the diffusion length is
equal to the film thickness, then about 80 % of the photogenerated charge will be
collected. Once the diffusion length reaches three times the film thickness, the charge
collection efficiency should approach unity [6.45]. It is apparent that the diffusion
Solid-state dye-sensitized solar cells 183

length can be increased by either improving the transport of charge carriers within the
device or by inhibiting electron hole recombination, and hence enhancing the charge
lifetime. Prior to improving something though, we need to know whether it requires
improvement. It is hence essential to accurately estimate the diffusion length in order
to direct research efforts.
The diffusion length for the solid-state DSC was first investigated by Kruger
et al. who performed intensity-modulated photocurrent and photovoltage spectros­
copy in order to derive the relative charge transport and recombination time constants
[6.46]. Intensity modulated photocurrent spectroscopy (IMPS) is a technique measur­
ing the periodic photocurrent response at short-circuit to a sinusoidal perturbation in
the light intensity superimposed upon a large bias background illumination. Intensity
modulated photovoltage spectroscopy (IMVS) determines the periodic response of
the photovoltage measured at open-circuit to the same modulated light source. We
will not go into precise details here, only mention that IMPS [6.47, 6.48] provides
information of the charge collection and recombination under short-circuit condi­
tions and IMVS [6.49, 6.50] gives data on exclusively the recombination, under open-
circuit conditions. Details can be found in the cited literature. Kruger at al. calculated
a diffusion length of 4.4 μιη for the solid-state DSCs studied [6.46]. This was longer
than the measured thickness of the cells (2 μιη) but suggested that improvements are
required if the cells are to function with thicknesses up 10 μιη, which is required for
optimum operation of the liquid electrolyte-based DSC. However, Kruger et al. made
an oversight of how the conditions in the solar cell vary with the applied bias: Under
short-circuit conditions, at which the transport was measured, the charge density was
low since a large amount of photo-generated charge could flow out of the device.
Under open-circuit conditions, where the electron lifetime was estimated, the charge
density was high since no charge could flow out of the device. Therefore, the cal­
culation of a diffusion length by comparing transport characteristics at short-circuit
and recombination characteristics at open-circuit has no direct relevance to any real
operating conditions. This has since been recognized and in a reprisal of the electron
diffusion length in solid-state DSCs, where the transport and recombination are esti­
mated under comparable conditions (constant quasi-Fermi Level for electrons in the
Ti0 2 ), Jennings et al. estimated a diffusion length of between 6 to 12 μιη for a similar
solid-state DSC to that measured by Kruger et al. [6.51].
Another technique for assessing the characteristic charge transport and
recombination rate constants consists in employing a transient photocurrent and
photovoltage decay method [6.5, 6.52]. This is in essence a real space version of
the IMPS and IMVS techniques, and should give exactly the same information. A
small light perturbation pulse (ideally temporally as short as possible) is superim­
posed upon a large background illumination. Following this pulse, the photovoltage
and photocurrent increase rapidly and then decay with a characteristic lifetime. If
the perturbation is kept small enough, the decays are close to mono-exponential.
From the voltage decay measurements, a pseudo first-order electron lifetime (r e )
can be estimated whereas an effective diffusion coefficient for electrons (De) can
be assessed from the current decay measurements. Here, De = d2/235 x rtrans, where
Ttrans is the current collection lifetime, and d is the film thickness. This relationship
between the current collection lifetime and diffusion coefficient assumes isotropic
184 Dye-Sensitized Solar Cells

three-dimensional transport [6.53]. For the current decay measurements, while the
charges are being collected, they also simultaneously recombine within the cell.
Therefore, the lifetime of the current signal (rsignal) is a combination of the lifetime
for the transport out of the cell (rtrans) and the recombination lifetime in the cell (r e ),
with the transport lifetime calculated as l/rtrans = l/rsignal - l / r e . The transient photo-
current measurements can be made at any given applied voltage by simply holding
a bias across the cell with a potentiostat, or suitably fast sourcemeter. The "voltage
perturbation" can also be measured at any given potential by simply holding the
corresponding current through the circuit with a galvinostat. Hence, under a given
background illumination, if we wish to study the cell under maximum power condi­
tions, the current corresponding to the max power point is held through the circuit
and the voltage perturbation recorded, after which the voltage corresponding to the
max power point is held across the device and the current perturbation recorded
[6.54], open-circuit is simply zero current with Voc and short-circuit is simply zero
volts with / sc .
The transient photovoltage and photocurrent decay technique has been used to
study the charge transport and recombination in solid-state DSCs over all working
conditions [6.55]. The calculated diffusion length versus the applied bias is given in
Figure 6.16 for a range of background light intensities. For the best solid-state DSCs,
the diffusion length is up to 20 μιη under short-circuit conditions and is relatively
invariant to changes in background light intensity. When an electrical bias is applied
against the cell to simulate working conditions, the diffusion length remains large
for bias potentials up to approximately 0.5 V, but then drops off rapidly as the bias
is increased towards an open-circuit potential. The main implications for the solar
cell design are that few further improvements to short-circuit current are possible,
whether by improving the transport or inhibiting the electron hole recombination.

Fig. 6.16 The electron diffusion length versus the applied bias measured over a range of light
intensities (see legend) for a solid-state DSC incorporating a K68 sensitizer. Figure drawn with
data from ref. [6.55].
Solid-state dye-sensitized solar cells 185

However, both the open-circuit voltage and the fill factor can be improved by either
of the above. We should note nevertheless that the diffusion length of 20 μιη was esti­
mated for solid-state DSC containing the best-performing sensitizer, which exhibits
significantly longer electron lifetimes (typically 2 times) than most other sensitizers
[6.8,6.19].
A third method by which charge transport and recombination can be studied
in dye-sensitized solar cells is impedance spectroscopy [6.38]. For this technique,
a small sinusoidal electronic potential perturbation is applied to the cell upon a
background-applied electronic bias, and the periodic current response of the cell is
recorded. The frequency of the perturbation potential is varied, and precise informa­
tion regarding the charge collection resistance at the electrodes, transport through
the Ti0 2 and hole-transporter and the back reaction (recombination) can be indi­
vidually derived since they respond in various frequency regions (have different
characteristic lifetimes). One of the derived parameters is the conductivity in the
Ti0 2 . For Ti0 2 , it is generally accepted that only electrons in the conduction band
are mobile and hence the conductivity is directly proportional to the population of
conduction band electrons. The population of these is related to the relative position
of the quasi Fermi level in the Ti0 2 (EFn) with respect to the conduction band energy
(ECB) according to:

(2)

where kB is Boltzmann's constant. Therefore, a comparison of other derived param­


eters at constant conductivity satisfies the requirement to compare parameters at a
constant quasi-Fermi level position [6.38].
Fabregat-Santiago et al. have contrasted the charge transport in solid-state
DSCs to that in electrolyte-based DSCs as a function of conductivity (E¥n position)
[6.38]. Interestingly, they found that the diffusion coefficient versus the conductivity
was close to identical for all devices, indicating that the electron transport was not sig­
nificantly influenced by the surrounding medium. The diffusion coefficients are plot­
ted as functions of the conductivity in Figure 6.17. The figure also indicates that the
electron transport generally limits the current collection and not the hole-transport,
in agreement with Kruger et al. and Snaith et al. [6.37, 6.46]. Less encouragingly
though, the recombination resistance is found to be considerably lower for the solid-
state cells, with the recombination resistance being proportional to the charge life­
time. This is also shown in Figure 6.17. These factors all combine to give a calculated
diffusion length of only 1 to 4 μιη. In this instance, it would therefore appear that
improving the diffusion length is an area that would wreak rewards. We note however
that most of the measurements were here carried out in the dark with an electronically
applied bias, where the conditions may differ significantly as compared to those under
standard solar cell operation.
Another interesting feature, elucidated by Fabregat-Santiago et al., was that
the resistance through the hole-transporter was relatively invariant to potential, only
reducing slightly when the potential was increased. Most significantly, at applied
potentials larger than 0.65 V, the resistance through the hole-transporter became
Next Page

186 Dye-Sensitized Solar Cells

Fig. 6.17 (a) Diffusion coefficients in several DSCs as functions of the electron conductivity
(quasi-Fermi level in the Ti02). (b) The recombination resistance normalized to volume in
several DSCs as a function of the electron conductivity, (c) The diffusion length in a solid-
state DSC as a function of the conductivity. These values have been calculated from modeling
impedance spectroscopy. Figure reproduced from an article in press, ref. [6.38], Copyright
2009 American Chemical Society.

greater that the resistance through the Ti0 2 , i.e., the systems appeared to switch
from electron-limited to hole-limited transport (assuming equal electron and hole
densities). The invariance of the hole conductivity to potential, combined with
the switching from electron-limited to hole-limited transport, could very well explain
the rapid drop in electron diffusion length at potentials over 0.5 V observed by Snaith
et al. (presented in Figure 6.16). The drop in diffusion length was due to the transport
lifetime not reducing as rapidly as the recombination lifetime with an increasing
potential. The titania has a charge-density-dependent mobility due to the increas­
ing population of conduction band electrons, whereas the spiro-OMeTAD has a
relatively invariant mobility with charge density. Hence, an invariance of the current
collection rate with increasing potential and light intensity (under the high potential
regime) was consistent with a transition from electron-limited to hole-limited current
collection.
CHAPTER 7

PACKAGING, SCALE-UP AND


COMMERCIALIZATION OF DYE SOLAR
CELLS
Hans Desilvestro, Michael Bertoz*, Sylvia Tulloch and Gavin Tulloch

7.1 INTRODUCTION

The Dye Solar Cell (DSC) technology continues to attract growing interest as a highly
credible alternative to both standard silicon photovoltaics and the more recently
developed thin-film technologies. Prices for poly crystalline silicon panels have not
decreased any further since mid 2004 and have plateaued just below US$ 5/Wp at
the retail level [7.1]. In Europe, poly silicon module prices are forecast to decline to
€ 2.30-2.45/Wp in 2009, while CdTe modules are expected to be at about € 1.55/Wp.
There is, of course, ongoing resistance among many green energy users to CdTe, and
the aggressive pricing is thus a reflection of that reluctance and the higher balance of
systems cost. The price for polysilicon is expected to be only temporarily low due to
the global financial crisis and UMG-Si (Upgraded Metallurgical Silicon) is not yet
considered viable. We note the comparison to the price touted by silicon PV industry
sources of US$ 40-80/kg to achieve competitive module prices. It is further noted
that China continues to produce mainly monocrystalline wafers of higher cost due
to capacity limitations. Nevertheless, the photovoltaic production has doubled every
two years, increasing by an average of 48 % each year since 2002, and rendering
it the world's fastest-growing energy technology. It is however apparent that, with
the exception of where there are subsidised feed-in tariff arrangements, 2009 prices
are too high for individual homeowners to install PV panels and to contribute at the
individual level towards a more sustainable future due mainly to system costs and
buy-back rates.

Ίη memory of our colleague and friend Michael Bertoz who tragically deceased in June 2009.
208 Dye-Sensitized Solar Cells

The present economic climate is not favorable for extending subsidies and gen­
erous feed-in tariffs any further (noting the 9 % reduction in feed-in tariffs in Germany
for 2009). With the expanding dissemination of PV installations, subsidies become an
increasingly large burden to governments and taxpayers, and are, as the 2008 example
of Spain shows, subject to a serious budgetary pressure. Moreover, standard silicon
technologies may not present the optimum solution for most inhabited regions, e.g.,
areas in higher latitudes with variable climates such as middle Europe and the UK
and even for the tropics with hazy and often overcast skies. In certain applications
such as on façades, traditional PV panels perform poorly due to the angle of incidence
being far from optimal. Silicon solar panels suffer from significant loss of efficiency
at lower light levels as a result of an increased electron-hole pair recombination due
to their being minority carrier devices.
In contrast, dye solar cells (DSC) represent majority carrier devices, and perform
particularly well under conditions corresponding to -0.2 to 0.5 AMI.5. Additionally,
they offer a whole range of other advantages as compared to traditional Si photovolta-
ics. DSCs display a superior dependence of performance on the angle of incidence
and their efficiency increases under diffuse light illumination [7.2] as well as at lower
light levels down to around 0.2 sun or lower.
DSC performance is much less dependent on panel temperatures than silicon.
Figure 7.1 shows IV curves for a typical commercial DSC design at temperatures
varying from -10 °C to 70 °C. The effects of temperature on some of the key char­
acteristics are summarized in Figure 7.2. While the open circuit voltage, Voc (O),
decreases linearly by ~2 mV/decade, which is similar to Si photovoltaic cells, the
maximum power point voltage of DSCs ( · in Fig. 7.2) remains, in contrast to silicon,
remarkably constant, varying by a mere ± 20 mV over the temperature range from
-10 °C to 70 °C (Fig. 7.3, left) and only decreasing slightly at temperatures above
45 °C. Fill factors, ff(A in Fig. 7.2), of DSCs increase as the temperature is raised and
only start to level off and subsequently decrease above 50-60 °C. The maximum jffas
a function of temperature depends on the light level, on the electrolyte composition
(specifically the I3 concentration), the electrolyte viscosity as well as other cell design
parameters such as anode-to-cathode spacing.
The lowered Voc at higher temperatures is largely compensated by a corre­
sponding increase in ff. Since /sc does not vary significantly with temperature, the
power Pmpp at the maximum power point (■ in Fig. 7.2) is fairly constant and varies
by only ± 12 % over the entire temperature range from -10 °C to 70 °C. Such a small
influence of the temperature on the power output from DSCs is in sharp contrast to
that of Si solar cells, which show a much more pronounced decrease in performance
[7.3] at higher temperatures, i.e., under real life sun light illumination. They also
display a much more pronounced drop in maximum power point voltage over a more
narrow temperature range A7(Fig. 7.3, right; note ArSi = 52 °C vs. ArDSC = 80 °C). A
power loss coefficient of 0.65 %/K [7.3] at Pmpp has been reported for Si-based solar
cells. Table 7.4 shows a comparison of power losses between DSC and Si for two
temperature ranges.
In addition, the DSC technology offers the potential of being considerably
more cost effective as opposed to traditional PV on the basis of the number of kWh
produced annually. The main reasons for the potentially lower cost include: a lower
Packaging, scale-up and commercialization of dye solar cells 209

Fig. 7.1 IV curves of metal-based flexible Dyesol DSCs from -10 °C to 70 °C at 1-sun
illumination.

Fig. 7.2 The influence of cell temperature on Voc (O), Vmpp (·),#(Δ), /mpp (♦) and Pmpp (■)
for metal-basedflexibleDyesol DSCs at 1-sun illumination.

embodied energy for DSC [7.4] as compared to other PV technologies [7.5], read­
ily available materials and significantly lower capital investments required for the
production equipment and facilities (e.g., no expensive clean rooms of Class 100
or lower).
210 Dye-Sensitized Solar Cells

Fig. 7.3 Left: power-voltage curves of metal-basedflexibleDyesol DSCs at a 1-sun illumina­


tion for temperatures ranging from -10 °C to +70 °C. The solid dots indicate maximum power
points. Right: Power-voltage curves of a single crystalline solar cell for temperatures ranging
from +28 °C to +80 °C [7.3]. The shaded areas indicate the spread of maximum power point
voltages.

Table 7.4 Comparison of P drop losses between DSC and Si for two temperature ranges.
Ui L
mpp

Temperature increase P mpp drop for DSC P mpp drop for c-Si

From 20 ° to 50 °C 5% 19.5 %
From 20 ° to 70 °C 15% 32.5 %

Other thin-film technologies, such as a-Si, CdTe or CIGS, still largely rely on
vacuum processes and therefore require costly equipment. In addition, they employ
toxic materials. For a-Si productions, significant amounts of extremely potent green­
house gases, such as PFCs (perfluorinated compounds) with atmospheric lifetimes of
thousands of years (CF4, SF6), are utilized [7.6], despite that there are a number of
development programs seeking to replace these with lower toxicity chemicals. CdTe
and CIGS thin-film products depend on relatively rare and expensive elements such as
Te [7.7], In [7.8] and Ga [7.9]. Freundlich's analysis, reviewing production and con­
sumption of key elements for various photovoltaic technologies, showed that known
In and Te reserves would not allow yearly CIGS and CdTe markets to grow to much
beyond the 10-GW level [7.10]. The availability of raw materials and costs for large-
scale DSC deployment is discussed in detail in section 7.4.1.
Despite the highly promising features of DSCs, no products are available in
2009 at a commercial scale. This study highlights key considerations for commercial
product development of DSC in terms of design (7.2.2), materials and their costs
(7.2.3, 7.41) and performance (7.2.4). Long-term product durability (7.3) under field
conditions is of utmost importance, particularly for building-integrated photovolta-
ics (BIPV) where a product lifetime of at least 25 years is required. While rapid
progress towards long-term stable modules and panels with attractive performance
Packaging, scale-up and commercialization of dye solar cells 211

characteristics is being made in many academic and industrial laboratories around the
world, large-scale commercialization poses additional challenges to sourcing produc­
tion quantities of specialized materials, which may not be required for any other appli­
cations, to set up high-throughput manufacturing and to assure profitable large-scale
commercialization. These topics are discussed in discussed in Sections 7.4 and 7.5.

7.2 FROM CELLS TO PANELS

7.2.1 Definitions
In the subsequent discussion, the terms 'cells', 'modules' and 'panels' are used as
follows:
Cells are units consisting of one anode-cathode system with a hermetic perim­
eter seal and two terminal contacts.
Modules are units with more than one anode-cathode system connected in
series or parallel, comprising cell-to-cell seals, often one perimeter seal and two ter­
minal contacts. Electrochemically, a single parallel connected unit (see P-design in
Fig. 7.5), is a 'cell'. But from its size and the complexity of current collectors, such a
unit is often referred to as a 'module'.
Panels are units comprising at least one module, an encapsulation structure
and two terminal contacts. A panel can consist of any number of cells and/or modules
connected in series and/or parallel. In addition, a panel may include electrical and
electronic control elements, such as protection and/or bypass diodes.

7.2.2 Designs
Since the 1991 seminal paper by O'Regan and Grätzel on the first assembly of a DSC
with > 7 % AM 1.5 efficiency [7.11], the design and the materials employed for single
cells have not changed drastically. Over the years, a number of module architectures
have been developed [7.12] in order to cover larger areas with series- and parallel-
connected cells, as well as to build up current and voltage. Each of the designs has its
distinct advantages and disadvantages in terms of material requirements, reliability,
manufacturing challenges and performance characteristics.
For single and poly crystalline silicon solar cells, the cell unit size is normally
no larger than 150 mm x 150 mm and is determined by the largest wafer size that
remains processable for module assembly without breakage (though there are indus­
trial goals to achieve larger wafers up to 250 mm x 250 mm). However, the paradox
is that it will be very difficult to move simultaneously towards thinner and thinner
wafers (for cost reasons) and to larger wafer sizes. Moreover, resistive losses for cur­
rent collection become more and more significant with larger and larger quasi-square
cells and thus require the more expensive (and light-blocking) silver. DSCs, on the
other hand, are not governed by the same limitations with regard to substrate size.
At least in one direction, the size can be considerably larger. Much more flexibility
is possible for DSCs in terms of module and cell size, which can relatively easily be
changed and optimized for any application. Moreover, cells and modules can be fully
212 Dye-Sensitized Solar Cells

Fig. 7.5 Cross sectional schematics of four main DSC module designs. In the parallel con­
nected design (P), current transport through the current collectors occurs perpendicular to the
viewing plane. With the other design current transport occurs primarily in the viewing plane.

integrated into large panels (> 1 m2) or continuously in a roll-to-roll process using
high speed and low cost manufacturing processes such as roller coating, printing and
lamination.
Parallel connected modules (P-design in Fig. 7.5) are inspired by the current
collection as used in standard silicon solar cells through a grid of highly conduc­
tive fine metal lines, usually silver or aluminum. In contrast to solid-state devices,
DSCs are photoelectrochemical systems, which normally contain organic solvent(s)
as well as the electrochemically corrosive iodine/iodide redox system. For this rea­
son, any internal current collection structures on the anode and cathode side need to
be thoroughly protected from corrosion (see 7.2.3). The P-design offers numerous
advantages, such as a higher performance for the same electrode width (see 7.2.4).
The cell units that need to be hermetically sealed are generally larger than with the
Packaging, scale-up and commercialization of dye solar cells 213

other designs, which in principle allows for a more compact design with a higher
percentage active area. Relatively large areas can be processed without the necessity
of extensive scribing of insulation lines. This can speed up the manufacturing process
and/or result in lower investments with regard to scribing equipment such as lasers.
On the other hand the P-design requires more conductive material, which can become
a significant cost factor (7.4.1). The P-design with low voltage/high current units is of
greater interest for large panels rather than small appliances.
The P-architecture is suitable for any glass/polymer/metal substrate combina­
tion as long as one substrate is transparent. For cases where both substrates are trans­
parent, the device efficiency may be further boosted by a back-scattering or reflection
layer.
Z-interconnected modules (Z-design in Fig. 7.5) offer the advantage of particu­
larly short cell-to-cell series connections. Thus, the interconnects can be based on less
conductive, corrosion-resistant metals [7.13]. The reliability of electrical contact to both
conductive substrates under all conditions of operation (i.e., extremes in temperatures)
is of utmost importance for this design. The Z-architecture provides a truly integrated
system to easily build up voltage. Cell-to-cell electrical insulation can be achieved
by removing small sections of the thin TCO l layer, e.g., through laser scribing. Cell-
to-cell seals need to be of high quality in order to exclude any electrolytic inter-cell
contact, otherwise electrolytic shunt currents occur, with ultimately I3 accumulating in
the positive terminal cell and fully or partially depleting in the other cells. The Z-type
interconnect system is in principle suitable for very large units. For the same electrode
widths, resistive current collection losses are considerably higher for the Z- as opposed
to the P-design (see 7.2.4), particularly for relatively short cells. Practical size limits
are dictated for glass-glass products by the flatness of the glass and the ability to accu­
rately reproduce cells. The Z-design is as suitable as the P-architecture for any glass/
polymer/metal substrate combination as long as one substrate is transparent. For metal
substrates however, an insulation layer is required in combination with a patterned con­
ductive layer on top of the insulation. For cases where both substrates are transparent,
the device efficiency may be further boosted by a back scattering or reflection layer.
W-interconnected modules (W-design in Fig. 7.5) appear attractive because of
the simplicity of the design which does not require any additional electrical intercon­
nect or current collection structures, thus potentially leading to lower costs. However, a
major disadvantage is that the cells are alternatively front and back illuminated and the
back illumination has a reduced output due to photon absorption by the electrolyte (tri-
iodide in particular). This can be compensated - at least for certain light conditions - by
making the back-illuminated cells wider in order to balance out for light absorption
through the electrolyte layer as well as for possible higher losses in electron collec­
tion from the titania layer. Nevertheless, it is practically impossible to design a system
where all cells are electrically well matched under all direct and diffuse light condi­
tions. Since each of the two substrates contains anode and cathode elements, optimum
processing conditions for each of the elements may be compromised. There is a higher
risk of cross contamination, and dye adsorbed on the counter electrode may lower the
performance. The cell-to-cell seals need to be of the same reliability as outlined for the

transparent conductive oxide.


214 Dye-Sensitized Solar Cells

Z-design. Resistive current collection losses are similar to Z-interconnected modules


(see 7.2.4). In principle, the W-design is suitable for any glass/polymer/metal substrate
combination as long as the front substrate is transparent. Since an effective, scalable
low-temperature Ti0 2 sintering process providing similar performance as -500 °C
processing of titania films has so far not been developed to be compatible with polymer
substrates, glass remains the preferred front substrate for this module architecture.
Monolithic modules (M-design in Fig. 7.5) are inspired by film and cell-to-cell
interconnect architectures employed for solid-state thin film photovoltaic panels such
as a-Si and CdTe. The monolithic design has been developed for DSC by Andreas
Kay [7.14] and patented by EPFL [7.15]. Only one conductive substrate is required
for this design, which can present a significant cost reduction of the components. The
porous, insulating separator layer has to be designed for an effective back scatter­
ing of non-adsorbed light after one passage through the Ti02-dye layer. The porous
counter electrode layer functions both as electrocatalyst for the I3 to Γ reduction and
as current collector. Critical areas for this design include the establishment of reliable
counter electrode-to-TCO electrical contacts and possibly a higher consumption of
costly dye due to adsorption on the high surface-area counter electrode. At the time of
writing, the substrate of choice for this design consists of TCO-coated glass. In con­
trast to the designs highlighted above, the monolithic modules cannot be made semi-
transparent. Their appearance will be much darker and there is virtually no freedom
for different colorations with this design.
Other module and panel designs are normally slight variations of the above or
combinations of single cells strung in series using different connectivity architectures.
As an example, G24i disclosed a design where single strip cells were attached to an
inert substrate through a thermally sensitive adhesive and externally interconnected
through a Z-type scheme [7.16]. A DSC panel assembly based on single cells may be
particularly attractive for consumer electronics applications, especially in combina­
tion with a high-speed production of relatively small units and high throughput pick-
and-place equipment on a suitable substrate.

7.2.3 Materials
Most materials used for single cells and DSC modules are the same, in particular:
substrates, semiconductor (nominally titania), dyes, electrocatalyst and electrolyte
systems. These materials, reviewed in other chapters of this book, are therefore not
discussed in detail here. Nevertheless, attention has to be paid to the different process­
ing conditions between manual processes for small laboratory cells and automatic
processing of modules and panels on high-speed lines, which can have implications on
product performance and reliability. Certain processes, such as resistive local heating
of the counter electrode TCO, easily implemented for small cells in order to achieve
reliable seals [7.17], cannot be readily scaled to » 100 cm2 modules, particularly in
the presence of current collectors, cell interconnects and electrical isolation patterns
previously applied to the substrates. Various sealing techniques may impose different
stress factors, particularly on the dye. Exposure to temperatures above 100 °C for
extended periods can lead to structural transformations of the dye-Ti02 system, result­
ing in less efficient electron injection and transport and thus a lowered device voltage
Packaging, scale-up and commercialization of dye solar cells 215

and fill factor. After processing, glass panels larger than -300 mm x 300 mm do not
necessarily remain flat at the ± 10 /mi scale. Such variations in cell thickness generally
affect cell fill factors and thus the electrical matching of cells in a module or panel.
The main materials that are expected to differ between laboratory cells and
larger devices for the commercial market include seals, cell-to-cell interconnects, cur­
rent collectors and busbars, panel encapsulants and, optionally, bypass diodes and
other protection diodes. Some of these components will be briefly reviewed in the
following paragraphs.
Seals have the function to hermetically enclose cells, modules and panels, and
more specifically, to:

• mechanically hold the two substrates together over the operating temperature
range;

• minimize solvent egress and gas ingress, particularly H 2 0 vapor;

• minimize leakage driven by capillary forces;

• prevent any electrolytic contact to corrodible current collectors and to other


cells.

Inadequate seal structures can lead to evaporation of solvent (even with gels)
and thus loss in conductivity and device performance. H 2 0 may accumulate inside
cells over extended periods of time, which may result in dye desorption and unwanted
chemical side reactions. Cell-to-cell electrolytic leak paths give rise to shunt cur­
rents, which lead to a serious electrochemically driven I3 redistribution between cells
of a module and ultimately to the complete loss of performance of such subunits.
Electrolytic contact to current collectors such as Ag collection fingers in parallel-
connected modules give rise to oxidation of Ag and thereby to an increase in series
resistance and, even worse, to the eventual complete depletion of I3 and thus perform­
ance. For instance, 100 mM of I3 in a 40-/mi thick electrolyte layer would be entirely
consumed by a corrosion current of 2.5 nA/cm2 over a year. Thus for a product life
of 20 years, average corrosion currents should not be much larger than 0.1 nA/cm2, a
level that has been shown to be achievable with common insulating films. Such cor­
rosion rates can be readily assessed through polarization curves or electrochemical
impedance spectroscopy.
Cell and module seals are normally in direct contact with the electrolyte system
and have to be chemically compatible with the solvent employed as well as all electro­
lyte components. Four main classes of sealing materials are under investigation:

• thermoplastic polymers

• elastomeric polymers

• adhesives

• glass frits
216 Dye-Sensitized Solar Cells

Thermoplastic gaskets based on 'hotmelts', such as the ionomer Surlyn® from


DuPont, have been successfully utilized to seal single cells for over ten years [7.18].
Silicone-based seals have been investigated for DSC applications for even longer
[7.19], due to their excellent chemical resistance characteristics. However silicones
can be rather porous and are generally poor vapor barriers. Consequently, a second­
ary seal may be required in order to provide adequate hermeticity. A large range of
adhesives, including epoxies, can be envisaged as the secondary seal. It is known, e.g.,
from plasma display panels, that glass frits provide the ultimate seal quality, and favo­
rable DSC stability results have been reported (see 7.3). Unfortunately the processing
of frit seals requires very high temperatures, normally in excess of 600 °C, for frits
that do not contain lead (frits need to be lead-free to avoid cell contamination [7.20,
7.21]. Subsequently, the dye solution has to be pumped into pre-sealed cells, which
is followed by the pumping through of a washing solution. Although glass frits are
suitable for display applications where the temperature extremes are small, the risk
of failure due to differential thermal stress in an operating DSC module is high as the
temperature range is above 100 °C. Finally, glass frits are expensive and inappropriate
for flexible products.
The same type of materials suitable for cell-to-cell and module seals can be
envisaged to protect current collectors from corrosion. Arakawa, et al. reported on the
successful protection of Ag current collectors in a 10 cm x 10 cm single cell module
of P-design by glass frit and polymer over-layers [7.22]. In 1995, STA (now Dyesol)
demonstrated that glass frits could be applied to protect grids, but that multiple appli­
cations were necessary to achieve pin-hole free coatings, and that the frits could not
be used with the rougher, higher-conductivity TCOs due to electrolyte leakage under
the glass-TCO seal.
Cell-to-cell interconnects for internal series connections are required for the
Z-design. Because of the short distance of 50 /mi or less across the cell, the intercon­
nect conductivity does not need to be very high and materials such as carbon, W, Ti
or metal oxides can be employed [7.13]. Some of these materials have the advan­
tage of showing virtually no corrosion in standard DSC electrolyte solutions. Suitable
interconnect materials can be used as particles embedded in a polymer matrix or as
narrow tapes, meshes, braids, wires, wires partially sheathed by insulation, etc. If
non-corrodible interconnect materials are utilized, requirements on cell-to-cell seals
become slightly less demanding. However, the prevention of electrolytic paths in
between cells still needs to be effectively suppressed.
Internal current collectors are imperative for the parallel design. Resistive losses
increase in close proportion to the square of the cell length. As a result of the limited cell
internal thickness available, the thickness, dcc, of the current collectors cannot be much
higher than 10-15 /mi. In order to maximize the device active surface area, they should be
as narrow as possible, with a width, wcc, of less than 1 mm. In a simplistic approximation,
i.e., assuming a uniform current density, j , the maximum device voltage drop, AV, along
the distance x to the closest anode and cathode busbars can be estimated according to Eq.
(7.1), where pcc represents the specific resistance of the current collectors:

(7.1)
Packaging, scale-up and commercialization of dye solar cells 217

Figure 7.6 presents the voltage drop along a 1-mm-wide and 15-/mi-thick cur­
rent collector as a function of the distance (JC) to the top and bottom busbars of a
parallel-connected cell consisting of 12-mm-wide T i 0 2 strips (schematic in insert).
It is assumed that there were equal current collectors on the anode and cathode sides
and that the current produced (e.g., at the maximum power point) was 15 mA/cm 2 .
This simple analysis demonstrates that for a decent AMI.5 performance and module
lengths (L = 2x) approaching 300 mm, current collectors were required with a specific
resistance of no more than 5 x 10~8 Ohm m, which is only approximately threefold
higher than for bulk silver. In reality, the current distribution would be rather inho-
mogeneous and only a numerical simulation can provide an accurate description (see
7.2.4). It should be noted that there have been recent developments in nano-scaled
conductors promising an increased performance - but at a price.
Panel encapsulants are required to:

• mechanically protect all components and electrical connections;

• not delaminate from cover sheets;

• protect all busbars and electrical contacts from atmospheric corrosion;

• protect all components from environmental stress factors (humidity, UV radia­


tion, hail, etc).

Fig. 7.6 Total voltage drop (i.e. from anode and cathode plates) along 1-mm-wide and 15-/mi-
thick current collectors for three specific resistance p cc values and as a function of the distance
x to the top and bottom busbars of a parallel-connected cell consisting of 12-mm wide Ti0 2
strips (schematic in insert) with equal current collectors on the anode and cathode sides. The
current density of 15 mA/cm2 was assumed to be homogeneous. The dotted lines in the insert
exemplify current collection paths.
218 Dye-Sensitized Solar Cells

Important requirements include a mechanical resilience, high optical transpar­


ency, affordability and, particularly for building-integrated applications, an extremely
good long-term stability without significant hazing or browning. For certain building-
integrated products, encapsulation based on toughened glass for the front sheet, in
combination with a transparent polymeric filler, can provide a reliable solution. A key
consideration is to avoid effects of thermal expansion and contraction of the poly­
meric layer over the operating temperature range. Edge seal and framing solutions
can be adapted from the double glazed unit (DGU) windows industry, however, the
disadvantage of such an encapsulation system is its weight.
In 2006, Dyesol and Corns began a worldwide collaborative effort to develop
light-weight flexible DSC panels to be integrated into roof structures, for example for
commercial buildings and warehouses. One of the main challenges was the develop­
ment of a reliable and cost-effective top polymer layer, capable of providing all the
functions described above. In addition, the top polymer layer and/or the transpar­
ent module substrate have to ensure excellent barrier properties in order to minimize
ingress of H 2 0 vapor and, if applicable, evaporation of solvent. Enormous progress
has been made with polymer-based transparent barrier materials over the last years,
mainly spurred by requirements from the OLED industry, where water vapor trans­
mission rates as low as 10~6 g/m2/day are demanded [7.23]. Estimates by Dyesol show
that DSC requirements are far less stringent, but still more demanding than those
imposed for see-through food packaging and pharmaceuticals. The cost/performance
trade-off is an important final design consideration. For outdoor applications, the DSC
technology should fulfill the requirements set by industrial standards such as IEC
61646 "Thin-film terrestrial photovoltaic (PV) modules".

7.2.4 Module performance - experiment vs. modeling


Due to the ongoing worldwide interest in DSC development, the performance of small
dye solar cells and the limiting factors are now very well understood [7.24]. The elec­
trochemical nature of DSCs and nonlinear, highly potential and temperature-dependent
processes governing dye solar cells, e.g., interfacial charge transfer, electron and ion
diffusion, partially in a tortuous porous network, render purely physical models rather
complex, with many parameters and assumptions governing them [7.25, 7.26]. On the
other hand, once the performance of a representative unit cell is known, extrapolation
to larger modules based on the same chemical components, film and cell thicknesses
is mainly controlled by straightforward electrical principles such as Ohm's law and the
Kirchhoff rules. The first step in such semi-empirical models is to correctly predict IV
curves of larger cells based on performance characteristics of smaller units (e.g., active
areas < 1 cm2). In the subsequent discussion, IV curves from Dyesol standard test cells
of a 0.88-cm2 active area were used as model input to simulate the performance of much
larger cells with the following current collection geometries and substrate combinations
(see Figs. 7.5 and 7.7, top and middle row):
• Z(FTO-ITO): glass-FTO2 (1 busbar on long side)/polymer ITO2 (1 busbar on
long side);
2
FTO: Fluorine doped tin oxide, ITO: Indium tin oxide.
Packaging, scale-up and commercialization of dye solar cells 219

• Z(Metal-ITO): metal (1 busbar on long side)/polymer ITO (1 busbar on long


side);

• P(Metal-ITO): metal (1 contact on short side), polymer ITO (2 busbars on long


sides).

Cells with two active area widths were assembled for each of the three types, and
representative samples can be seen in Figure 7.7 (middle row). The goal was then to
simulate the electrical influence of cell width and design on IV characteristics for TCO-
based substrates with a typical sheet resistance of 15 Ohm/square compared to metal sub­
strates of very high conductivity. For this reason, all cells, including the 0.88 cm2 test cell
to provide the empirical IV input into the model, were illuminated in reverse mode, i.e.,
from the counter electrode side. TV curves of the test cell determined at various illumina­
tion levels were corrected by the series resistance Rs, determined from electrochemical
impedance spectroscopy, to provide /^-corrected 7V0 curves with V0 given by Eq. (7.2):

(7.2)

2-D electrical models were set up in COMSOL Multiphysics® (AC/DC mod­


ule) to emulate the three cell models shown in Fig. 7.7 (top). The active areas acted

Fig. 7.7 Top row: Cross-sectional schematics of three types of cell configurations. Middle
row: Photos of cells assembled according to the three configurations with two widths per con­
figuration. From left to right: Z-type glass-FTO (Ti02)/polymer-ITO (Pt); Z-type metal (Ti02)/
polymer-ITO (Pt); P-design metal (Ti02)/polymer-ITO (Pt). Bottom row: Calculated current
density maps at the maximum power point at 1-sun illumination for the 6-cell configurations
and geometries shown in the middle row.
220 Dye-Sensitized Solar Cells

as a current source and drain, governed by area-normalized IV0 functions for varying
light levels while the current collection from active and inactive areas was character­
ized by device geometry and the electrical conductivity of various sections of the
device. The lines in Figure 7.8 depict the simulated device performance for the three
models as a function of the active area width. As can be seen, a higher substrate
conductivity and narrower Ti0 2 widths clearly led to higher performances. As shown
in Figure 7.9, the current paths with the P-design were shorter than those with the
Z-architecture. The performance sequence (7.3) may appear to be trivial, but compari­
sons to experimental data demonstrate how well the relatively simple model describes
the actual performance. The drop-off in performance with increasing cell width was
much more pronounced for the Z-type current collection geometry.
(7.3)

Current density maps determined through numerical simulation showed that


the assumption of a uniform current distribution was only justified for some of the
cases. Particularly with larger Ti0 2 cell widths and the Z-design based on substrates
of varying sheet resistance, the current distribution became very inhomogeneous as
qualitatively expected from Figure 7.9 (top). For the Z(Metal-ITO) configuration, i.e.,
with metal as the anode substrate, the lower resistance current paths visualized in
black in Figure 7.9 (top) are preferred over the current paths shown in gray. This was
numerically confirmed by the current distribution maps shown in Figure 7.7 (center
of the bottom row). The numerical model quantitatively demonstrated the higher cur­
rent density in P-connected cells close to the busbars compared to the center of the
active area (Fig. 7.7, bottom row, right). Qualitatively, such a current distribution was

Fig. 7.8 The performance at 1-sun illumination for three designs according to Figure 7.7 as a
function of Ti02 electrode width. The solid lines were calculated from numerical simulation
and the points correspond to experimental values.
Packaging, scale-up and commercialization of dye solar cells 221

Fig. 7.9 A schematic of current lines in cells based on the Z- (top) or the P-design (bottom).

expected since the current along the black arrows in Figure 7.9 (bottom) experienced
less Ohmic resistance than along the light gray arrows.
Such calculations are very useful for optimizing the cell geometry for dif­
ferent applications. Practically, numerical simulation becomes even more useful
when whole modules or panels can be quantitatively assessed. While it is obvious
that more narrow cells provide an enhanced performance per active area, seal areas
become proportionally more significant with decreasing cell widths, thus decreas­
ing the proportion of active area in a panel and possibly reducing the performance
per panel area. Customers are only interested in the latter and they are even more
interested in the lowest possible cost for an acceptable performance. A complete
panel model has to include all current collectors, busbars, module-to-module elec­
trical contacts and any additional electronic elements such as bypass and/or panel
protection diodes. Such models will be much more involved and have to be con­
structed in a modular way since there can quickly be hundreds, if not thousands,
of boundary conditions for the model of a complete panel. Sophisticated numerical
modeling will however provide the ultimate development tool for optimizing per­
formance and current collection paths, as well as minimizing cost, e.g., the amount
of silver employed.
Figure 7.10 shows simulated voltage (top) and current density maps (bottom)
for the respective maximum power point at l-(left) and 0.33-(right) sun illumination
intensities for a parallel-connected 188 x 92 mm2 module with 6 Ti0 2 stripes on FTO,
a counter electrode with metallic conductivity, Ag current collectors and one higher
conductivity busbar on the left side of the anode. The illumination was assumed to
222 Dye-Sensitized Solar Cells

Fig. 7.10 Voltage (top) and current density (bottom) maps calculated for the respective max­
imum power point at 1-sun (left) and 0.33-sun (right) illumination for a parallel-connected
188 x 92 mm2 module with six Ti02 stripes on an FTO-based anode, Ag current collectors and
one higher conductivity busbar on the left side of the anode and a counter electrode (cathode)
with metallic conductivity.

come from the Ti0 2 side. Data was calculated based on a high performance 0.88 cm2
test cell comprised of two FTO substrates connected in Z-configuration with front-
illumination (see dotted curves in Fig. 7.11). Figure 7.10 graphically quantifies
the voltage losses in various sections of the module. At high illumination densities
(Fig 7.10, top left), the main voltage drops occurred within the active areas, along the
main busbars and, particularly for the left-most Ti0 2 strip, also along the current col­
lection fingers. The voltage distribution at the maximum power point for 0.33 sun was
much more uniform (see Fig. 7.10, top right). At 1 sun, the current density variation
ranged from 100 % to 67 % while the corresponding variation at 0.33 sun was only
between 100 % and 91 % (see Fig. 7.10, bottom left and right).
The calculated TV curves for the module are represented by the solid curves in
Figure 7.11 for three illumination levels and based on the device active area. As can
be seen, it is possible to design modules that are as good or even better than standard
test cells at least when assessed based on only the active area. In fact, Dyesol standard
test cells are deliberately designed to be a relevant small-scale model for perform­
ance of commercially viable products. It would be easy to boost the performance
of Dyesol standard test cells through 'tricks' such as more effective circular current
collectors. However, such features could not be transposed to larger area modules and
would therefore be an artifice. Table 7.12 shows the key performance characteristics
of a Dyesol test cell vs the simulated 1-cell module based on the active and total area.
The simulated design indicates module efficiencies of 6-7 % at < 0.5 sun light levels,
Packaging, scale-up and commercialization of dye solar cells 223

Fig. 7.11 JV curves for the module shown in Figure 7.10 for three illumination levels and
based on the device active area (solid curves) vs. JV curves of a 0.88 cm2 Dyesol test cell con­
nected in Z-configuration.

Table 7.12 Some key parameters of Dyesol test cells in comparison with the simulated per­
formance of the parallel-connected module, see also Figures 7.10 and 7.11.

Dyesol test cell 6 strip P-module, 6 strip P-module,


based on active area based on total area

1 0.33 0.1 1 0.33 0.1 1 0.33 0.1


sun sun sun sun sun sun sun sun sun

Efficiency (%) 6.04 8.07 9.09 6.55 8.30 9.23 4.85 6.15 6.84
^mpf ,(V) 0.49 0.56 0.55 0.51 0.56 0.55 0.51 0.56 0.55
2
^mpp (mA/cm ) 12.40 4.77 1.66 12.85 4.89 1.68 9.52 3.62 1.24
ff 0.48 0.66 0.73 0.52 0.68 0.75 0.52 0.68 0.75

, ΛηΡΡ = Voltagis, current density at inaximum power poiitit.

based on the total module area. Table 7.12 further demonstrates that the maximum
power point voltage Vmpp remains rather constant over a wide range of light levels.
This is a typical feature of dye solar cells and it provides a significant advantage over
more traditional solid-state photovoltaic devices.
In summary, numerical simulation can be used as a tool to thoroughly under­
stand the relationship between performance and module design, current collection
224 Dye-Sensitized Solar Cells

geometries and the cross section of current collectors of a given electric conductivity,
as well as to optimize a design for virtually any practical application.

7.3 LONG-TERM STABILITY - THE KEY TO INDUSTRIAL


SUCCESS

Assuring long-term stability and product robustness is a very important aspect of the
DSC scale-up process, especially for building-integrated applications. While DSCs
based on volatile solvents such as acetonitrile or methoxyacetonitrile, used for high­
est-performance cells, show poor long term stabilities, there is nowadays an avail­
ability of numerous electrolyte systems that are based on low-volatility solvents with
boiling points of > 150 °C or even > 200 °C. Also completely solvent-free electro­
lyte systems based on room-temperature ionic liquids (RTIL) have been developed
[7.27]. The DSC performance of RTIL-based electrolyte systems is lower, i.e., 8.2 %
at AM 1.5 for the best small laboratory cells reported thus far [7.28]. This can be com­
pared to DSCs containing volatile solvents, where the 2009 world record (based on
Dyesol Ti0 2 paste) was at 12.2 % under AM 1.5 irradiation [7.29]. In addition, ionic
liquids are generally still considerably more costly [7.30] than electrolyte systems
based on solvents such as nitriles or lactones.
To achieve an acceptable product lifetime, a durability has to be achieved
at four levels, i.e., at the molecular, cell, module and finally at the system level.
A. Hagfeldt pointed out that the "Ti0 2 film on its own does not conduct electrical
current, the dye on its own cannot be exposed to sunlight and the electrolyte is cor­
rosive" [7.31]. However, if combined all together in the correct balance, a stable and
effective solar cell is obtained, thanks to the chemically defined, kinetically control­
led interplay of the components in the presence of light and charge-transferring elec­
trodes. A system's robustness at the molecular level was estimated by M. Grätzel,
through an analysis of the kinetics of electron injection, dye regeneration and chemi­
cal and photochemical side reactions for a standard Ru-dye based cell chemistry. It
was concluded that more than 100 million sensitizer turnovers 3 are feasible, cor­
responding to a lifetime of 20 years under average solar illumination [7.32]. The
stability at the levels of increasing integration and complexity is discussed in the
following sections.

7.3.1 Single cells


Apart from the electrolyte formulation and a judicious choice of sensitizer, hermetic
seals are the key to DSC longevity. The functions of the seals have been reviewed in
Section 7.2.3. Since DSCs experience temperatures up to 70-80 °C under full sun
illumination for up to 10 % of their operating life, a long-term assessment needs to
be undertaken at test temperatures well above 50 °C A number of laboratories have

3
One turnover corresponds to one complete cycle of dye excitation, electron injection into
the titania conduction band and dye regeneration through I~.
Packaging, scale-up and commercialization of dye solar cells 225

demonstrated stability of DSC output over 1,000 hours at 60 °C, with most of the
systems being based on various Ru-bpy dyes [7.33-7.37] and, very recently, even with
metal-free coumarin dyes at 50-55 °C [7.38]. Hinsch et al. [7.21] showed that thermal
stress due to 1,400 h storage at 85 °C in the dark leads to a more marked loss of per­
formance than 8,000 h of illumination at a constant 20 °C cell temperature under light
levels corresponding to 2.5 suns [7.39]. Pettersson et al. observed a good stability for
monolithic single cells under relatively low light level illumination (5,000 lux) over
180 days (-4,300 hours) [7.40]. The authors further demonstrated the importance of
UV-light filtering (< 400 nm) in order to avoid Ti0 2 band gap illumination. Yanagida
et al. observed a markedly improved stability at 85 °C in the dark through addition of
an organogelator to a RTIL-based electrolyte system. While a DSC with solvent-free
un-gelled electrolyte lost close to 30 % of its performance over 1,000 h at 85 °C in the
dark, the addition of the organogelator was reported to thoroughly stabilize the DSC
efficiency under light [7.41].
Dyesol, the largest industrial player in the DSC field, has established exten­
sive expertise and understanding of long-term durability at the cell and the module
level through a range of accelerated tests, such as continuous exposure to artificial
sun light at elevated temperatures, high temperature storage, thermal cycling and
humidity tests. A promising stability over 20,000 hours of simulated sun light at an
average light level corresponding to 0.8 sun has been presented [7.42], particularly
for cells based on the hydrophobic ruthenium-dye Z907. Meanwhile the same cells
have accumulated 27,500 hours of light soaking. Continuous illumination over such a
period of time is equivalent to an accumulated irradiation energy of 22,000 kWh/m2.
Experimentally, it was determined that the annual device temperature, assessed solely
from the number of hours of sunshine under a virtually cloudless sky, averaged about
45 °C when the devices on Dyesol's roof (Southern hemisphere) faced the North at
an angle of 35 °C from horizontal (=Canberra latitude). A common approach is to
assume that the rate of aging of polymer-based products or polymer-encapsulated
devices increases by a factor of 2-3 for a raise in temperature of 10 °C. Dyesol's accel­
erated aging tests thus indicated a service life of at least 40 years in areas experiencing
1,000 kWh/m2 annual solar irradiation (e.g., Middle Europe or Southern England) and
25 years in areas such as Southern Europe.
More recently, Dyesol has further optimized its cell chemistry, including
electrolytes, Ti0 2 pastes and films (2008 Dyesol Technology). Data over close to
11,000 hours of accelerated solar exposure of improved cells is shown in Figure 7.13
for DSCs incorporating N719- and Z907-based cells and an electrolyte solution con­
taining a low volatility solvent. Note that the same type of Ti0 2 , i.e., Dyesol Ti0 2
Paste DSL 18NR-T, enabled the group of Prof. M. Grätzel to increase the DSC world
record for single cells to 12.2 % at AM1.5 [7.29].
Even with the hydrophilic and thus inherently less long-term stable N719 sen-
sitizer (higher tendency towards desorption from Ti0 2 surface following moisture
ingress), a remarkable long-term stability has thus far been achieved in accelerated
aging tests with cells based on the 2008 Dyesol Technology. Using the same extrapola­
tion method as above, this result already corresponds to close to 17 years of service in
areas experiencing annual solar irradiation of 1,000 kWh (e.g., Middle Europe). The
main reasons for some loss in cell efficiency include the open cell voltage decreasing
226 Dye-Sensitized Solar Cells

Fig. 7.13 AM1.5 ("1 sun") and 0.33 AM1.5 ("0.33 sun") efficiencies of N719 (--) and Z907
(-) based cells as a function of the light soaking time. Active area: 8x11 mm2. The cells were
held close to the maximum power point and illuminated at an average light level of 0.8 sun
resulting in a cell temperature of 55-60 °C. The cells were periodically characterized at ambient
temperature through their IV curves at varying light levels.

over the first 1,000 hours of the light soaking test and a slight decrease in fill factor
beyond 6,000 hours.
In order to better understand the initial lowering of open circuit voltages and
some decrease in fill factor, electrochemical impedance spectroscopy (EIS) was per­
formed. EIS has proven to be a very powerful technique to distinguish resistive and
capacitive cell components and interfaces, along with charge transfer and electron
and ion diffusion processes [7.43, 7.44]. All data from fresh as well as aged cells
was fitted using the simple equivalent circuit shown in the insert in Figure 7.14(a). In
principle, EIS should be measured in a 3-electrode configuration including a reference
electrode. On the other hand, measurements at a constant cell voltage will, depending
on the current drawn, polarize the counter and the Ti0 2 electrode differently. In order
to measure at a quasi-constant Ti0 2 potential without the experimental complication
of inserting a reference electrode into each cell, the DC bias E was adjusted for each
measurement to fulfil Eq. (7.4) within ±5 mV or better. Here, j represents the dark
current density and Rs and Rct the cell series resistance and the counter electrode
charge transfer resistance, respectively.
(7.4)

Figure 7.14(a) shows that the electron back transfer reaction resistance, Rhr,
decreases significantly over the first 1,000 hours of illumination due to electrochemical
reaction (7.5). The chemical capacitance, Ccc on the other hand, initially increases rela­
tively quickly and then more slowly with the illumination time. As a result, the drop in
open circuit voltage can clearly be ascribed to an increased rate of reaction (7.5) which,
as evidenced by the concomitant increase of Ccc, is due to a shift of the Ti0 2 conduction
Packaging, scale-up and commercialization of dye solar cells 227

Fig. 7.14 EIS parameters as functions of the light-soaking time determined in parallel to the
cell performances shown in Figure 7.13. (a) Ti0 2 electrode, (b) counter electrode, (c) series
resistance. The parameters were fitted according to the equivalent circuit shown in the insert.
The two RC elements were fitted by constant phase elements. In order to achieve an improved
consistency between measurements for different cells and irradiation times, the alpha param­
eter for the counter electrode was held constant at 0.85.
228 Dye-Sensitized Solar Cells

band in the positive direction, (i.e., closer to the I3/Γ potential). Such a shift indicates
that the Ti0 2 surface charge becomes more positive over the initial 1,000 hours of light
soaking, possibly as a result of increased surface protonation levels.

(7.5)

On the other hand, EIS data shows that the counter electrode performance,
based on Dyesol Platinum Paste PT1, is definitely not a contributor to any degradation
in cell performance, even over very long service times. On the contrary, the counter
electrode electrokinetics became even faster over time as shown by the decrease in
charge transfer resistance, RcU in Figure 7.14(b) while Cdl, the platinum double layer
capacitance, remained virtually constant. This shows that there was no loss or passi­
vation of the electrochemically active area and that no significant loss of I3 occurred
during the course of the accelerated long-term tests. The latter conclusion was further
corroborated by a virtually constant Nernst impedance (not shown in Fig. 7.14), due
to I3 diffusion in the electrolyte layer.
EIS revealed that the series resistance, Rs, increased noticeably beyond 4,000-
6,000 hours, which was the most likely reason for the decrease infillfactor, particularly
at the higher sun levels. It is thus important to maintain the electrical leads-to-TCO
Ohmic contacts low and stable over prolonged times of usage, in order to maintain a
stable fill factor and performance.

7.3.2 Modules
Increased cell and module areas create greater materials engineering challenges due
to the fact that hermetic sealing is more demanding for larger units than for small
laboratory cells. For DSC products, where both substrates are glass-based, the lack
of flatness of commercially available glass at the ±10 /mi scale does not permit single
modules to be much larger than -300 x 300 mm2. Electrical series and parallel con­
nections need to provide low contact resistance and the current collection resistance
should be low and independent of any thermal stress imposed by continuous day/
night cycles. Silver, copper and aluminum, the major candidates for current collec­
tion from > 100 cm2 areas based on electrical conductivity and cost, corrode in the
presence of standard DSC electrolyte systems and therefore need to be thoroughly
protected. Electrolytic cell-to-cell leakage has to be avoided for reasons explained in
7.2.3. Moreover, cell-to-cell variability should be minimized in order to avoid module
imbalance, which can lead, in the worst case, to reverse polarity operation of lower
performing cells and to module damage. Nevertheless, in contrast to standard p-n
junction devices, DSC is much more resilient to temporal natural shading. However,
for applications where regular hard shading, e.g., shades cast by a building, occurs on
a daily basis, diode protection may be required and should be evaluated on a case-by-
case basis. On the other hand, there is now sufficient expertise from numerous indus­
trial and academic laboratories which indicate that, apart from sufficiently hermetic
polymer substrates for flexible modules and possibly long-term corrosion protection,
no materials development breakthrough is required to achieve excellent DSC service
life at the module level, provided that prolonged continuous temperatures above 70 °C
Packaging, scale-up and commercialization of dye solar cells 229

can be avoided. There are many applications, in moderate climates or indoors, or even
on façades in tropical climates, where such conditions are met. Development work is
however ongoing to extend long-term stability to 80-85 °C.
Compared to the cell level, there is less stability data available for DSC modules.
Recently, Arakawa et al. presented a promising stability performance for 10 x 10 cm2
multi-strip single cell modules with current-collecting Ag grids [7.22]. Cells using
the 'black dye' [7.45] and being filled with an MPN 4-based electrolyte system were
subjected to thermal cycles between -40 °C and +90 °C according to Japan Industrial
Standard C-8938. Two hundred such cycles resulted in an efficiency degradation from
about 6.5 % to below 4 %, i.e., in a relative performance loss of around 40 %. In com­
parison, virtually no loss of performance was recorded after 10 heat-humidity cycles
(-40 °C to +90 °C at 85 % humidity), demonstrating that prolonged heat stress, as
opposed to high humidity and short bursts of high heat, is more severe to well-sealed
DSCs using the black dye. Separate observations by Dyesol have indicated that the
black dye may be susceptible to detrimental processes involving the dye and/or the
titania/dye interface at elevated temperatures.
Toyota Central R&D Laboratories and Aisin-Seiki have reported that a 64-cell
module [7.46] under outdoor conditions during half a year lost around 15 % of its
performance, which was ascribed to "poor sealing" [7.47]. Very recently, the same
group presented outdoor stability data over 2.5 years of N719 and γ-butyrolactone-
based 10 x 11 cm2 modules consisting of 3-cell monolithically series-connected sub-
modules, where three such submodules were connected in parallel [7.47]. The authors
noticed the same type of performance degradation patterns as with single cells illumi­
nated under simulated solar light at 60 °C for 1,000 h, in other words:
• initially a rather rapid decrease of open circuit voltage followed by a more
gradual decrease in Voc;
• a concomitant increase of dark currents;
• initially a slight increase of short-circuit currents followed by a very slight
decrease over extended periods of time;
• a continuous gradual drop in fill factor.
These long-term effects on the device performance are very similar to those
observed by Dyesol and described Section 7.3.1. In the same study [7.47], Toyota
and Aisin-Seiki compared EIS and Raman data for fresh and aged cells. While the
main vibrational features of the dye remained unaltered, a certain decrease in I3 was
evidenced by Raman spectroscopy and the authors rationalized the increased Nernst
diffusion impedance by a lowered I3 concentration in aged cells. However, the Dyesol
cells represented in Figure 7.14 showed no evidence of an increase in Nernst imped­
ance as a function of time. EIS data presented by Toyota and Aisin-Seiki researchers
was obtained under open circuit conditions under illumination corresponding to a
0.7-sun light intensity [7.48]. Thus their EIS results differed considerably from those
shown in Figure 7.14. Moreover, they used porous carbon black-based counter elec­
trodes instead of Pt employed by Dyesol.
4
MPN = 3-methoxypropionitrile.
Next Page

230 Dye-Sensitized Solar Cells

Sastrawan et al. showed, with a 29-cell 30 x 30 cm2 module sealed by glass


frits, but not filled with any electrolyte solution, that the seals and Z-interconnects can
withstand 50 thermal cycles between -40 °C and +80 °C without giving rise to any
increase in interconnect resistance [7.20].
Dyesol semi-transparent modules consisting of a > 100 cm2 active area, which
are suitable as building blocks for solar window applications, were investigated with
regard to their long-term stability under the same conditions as those used for single
cells. Achieving the same high level of reliability, stability and efficiency as for sin­
gle cells remains challenging. However, solid progress has been made as shown by
Figure 7.15, where the power output after 4,000 hours of light soaking at -0.8 sun
and 55-60 °C remains almost identical to the initial performance at all sun levels
tested. It should nevertheless be pointed out that the initial performance of the module
shown in Figure 7.15 is lower when compared to Dyesol test cells and the recovery
over the first 1,000 hours is therefore larger for the module shown in Figure 7.15 than
for the test cells. Development work to improve initial module performance, device
resilience and reliability is presently underway and the new designs will be released
by Dyesol in 2009.

7.3.3 Panels
For practical applications, modules need to be assembled into full panels and mounted
onto or designed into building roofs and façades, or integrated into windows or

Fig. 7.15 The power output at three light intensities of an N719-based module (photo) as a
function of the light soaking time. The module was held close to the maximum power point and
illuminated at an average light level of 0.8 sun, resulting in a module temperature of 55-60 °C.
CHAPTER 8

HOW TO MAKE HIGH-EFFICIENCY


DYE-SENSITIZED SOLAR CELLS
Seigo Ito

In this chapter, the optimization of the fabrication technologies for the high-efficiency
dye-sensitized solar cells (DSC) are discussed. Techniques include TiCl4 treatments
for the photoelectrodes, and the application of a transparent nanocrystalline-Ti02-
layer, a light-scattering layer and an anti-reflecting film on TCO-substrates. The TiCl4
treatments are necessary to improve the mechanical strength of the Ti0 2 layer. The
thickness of Ti0 2 layer affects the photocurrent and the photovoltage of the devices.
Furthermore, the photocurrent can also increase with introduction of an anti-reflect­
ing film. These components have significant influences on the energy-conversion
efficiency.

8.1 INTRODUCTION

Dye-sensitized solar cells (DSCs) have recently undergone intensive investigation as


a promising inexpensive alternative to conventional p-n junction solar cells [8.1-8.4].
The use of dyes and nanocrystalline Ti0 2 is one of the most promising approaches
towards the realization of both high performance and low cost, thanks to their low
material cost and ease of chemical manufacturing [8.5]. A high light-to-electricity
conversion efficiency results from a large surface area of porous Ti0 2 electrodes, onto
which the dyes can be sufficiently adsorbed. In order to prepare the Ti0 2 electrodes
for high efficiency dye-sensitized solar cells (i.e., with conversion efficiencies in
excess of 10 %), we are using screen printing (for nanocrystalline- and submicron-
crystalline- Ti0 2 layers) and chemical-bath deposition (for TiCl4 treatment) [8.6]. In
addition, a photon-trapping effect created by the use of a system of transparent and
light-scattering layers (double layer) and an anti-reflecting film (ARF) have been used
to enhance the quantum efficiency as measured by the incident photon-to-electricity
conversion efficiency (IPCE).
In this chapter, the fabrication method and the influence of different procedures
on the photovoltaic performance for the high-efficiency DSC are illustrated.
252 Dye-Sensitized Solar Cells

8.2 EXPERIMENTAL CONSIDERATIONS

8.2.1 Preparation of screen-printing pastes


Two kinds of T i 0 2 pastes have been prepared, the first containing nanocrystalline-
T i 0 2 (20 nm) and the second macrocrystalline-Ti0 2 (400 nrn) particles; these lay­
ers provide the transparent and light-scattering layers, respectively [8.7]. Figure 8.1
shows the preparation scheme for the nanocrystalline-Ti0 2 (20 nm) paste. Two kinds
of ethyl cellulose (5-15 mPas at 5 % in toluene:ethanol/80:20 at 25 °C, #46070, Fluka;
30-50 mPas at 10 % in toluene:ethanol/80:20 at 25 °C, #46080, Fluka) were dissolved
first in an ethanol solution. Each concentration of the ethyl cellulose was controlled
to 10 wt%, yielding a total of 10 wt% of ethyl cellulose. Separately, 12 g (0.2 mol) of
acetic acid was mixed to 58.6 g (0.2 mol) of titanium iso-propoxide under stirring at
room temperature, drop by drop. Then the modified titanium iso-propoxide precursor
was stirred for 15 minutes and poured into 290 ml of water in a second under stirring
(700 rpm), resulting in a white precipitate.
After one-hour of stirring, 4 ml of 65 % nitric acid was added to the dispersion,
which was heated to 78 °C over a period of 40 minutes and maintained for 75 min­
utes. Then, the heating was stopped and water was added to the dispersion to a total
volume of 370 ml. The resultant dispersion was put in a 570-ml titanium autoclave
and heated at 250 °C for 12 hours. After that, 2.4 ml of 65 % nitric acid was added
and sonicated using a 200-W ultrasonic titanium probe with 30 cycles (2 seconds work
followed by 2 seconds rest). The resultant colloidal solution was concentrated with a
rotary-evaporator to obtain 13-15 % Ti0 2 . Finally, it was centrifuged and washed with
ethanol three times to remove nitric acid and water. After the three centrifuge cycles,

Fig. 8.1 Procedure for Ti0 2 paste for screen-printing [8.7].


How to make high-efficiency dye-sensitized solar cells 253

the precipitate contained 40 wt% Ti0 2 in ethanol with traces of water. Anhydrous
terpineol (Fluka) is added to the mixture solution of two types of ethyl cellulose
with ethanol to obtain the ratio: 16 g of Ti0 2 , 4.48 g (44.8 g solution in ethanol) of
ethyl cellulose (#46070, Fluka), 3.52 g (35.2 g solution in ethanol) of ethyl cellulose
(#46080, Fluka) and 56 g of terpineol and ethanol added to obtain 200 ml total.
The mixture is then stirred by using a magnet tip and sonicated with an ultra­
sonic horn (Sonics&Materials). The contents in dispersion were concentrated by using
an evaporator at 40 °C under 120 mbar initially. The pressure is reduced to 10 mbar
to remove ethanol and water. The pastes were finalized with a three-roll-mills grinder
(M-50, EXAKT, Germany, Fig. 8.2). At this point, the paste is prepared consisting
of 16.2 % of the 20-nm-sized Ti0 2 and 4.5 % ethyl cellulose in terpineol. Finally,
we obtain 80 g of paste containing 20 wt% Ti0 2 , 5.6 wt% ethyl cellulose (#46070,
Fluka), 4.4 wt% ethyl cellulose (#46080, Fluka) and 70 wt% terpineol.
For the paste used in the light-scattering layers, just after the autoclaving, the
20-nm Ti0 2 nanoparticles [8.1] are mixed with 400-nm Ti0 2 particles (CCIC, Japan).
The paste is composed of 28.6 % of 400-nm-sized Ti0 2 , 2.9 % of 20-nm-sized Ti0 2
and 7.2 % ethyl cellulose in terpineol. The other procedures were same with that of
20-nm Ti0 2 paste (Fig. 8.1).

8.2.2 Synthesis of Ru-dye


The synthesis of c/1s,-di(thiocyanato)-A/,A^/-bis(2,2/-bipyridyl-4-carboxylic acid-4'-
tetrabutylammonium carboxylate) ruthenium (II) (N-719) has been reported in a
recent paper [8.8]. The Chromatographie purification of N-719 is carried out three
times on a column of Sephadex LH-20 using the following procedure [8.9]. The N719
complex is dissolved in water containing two equivalents of tetrabutylammonium
hydroxide. The concentrated solution is then filtered through a sintered glass cru­
cible and charged onto a Sephadex LH-20 column, which is prepared in water. The
adsorbed complex is eluted using water. The main band is collected and the solution
pH is lowered to 4.3 using 0.02 M HN0 3 . The titration is carried out slowly over a

Fig. 8.2 Three-roll mill forfinishingthe Ti02-paste procedure: putting the paste between the
rolls (a) and taking out the paste in front of the mill (b).
254 Dye-Sensitized Solar Cells

period of three hours. Then, the solution is kept at - 2 0 °C for 15 hours. After allowing
the flask to warm to 25 °C, the precipitated complex is collected on a glass frit and air
dried. The same purification procedure was repeated three times to get pure N-bonded
isomer complex.

8.2.3 Porous-Ti02 electrodes


Figure 8.3 shows the preparation scheme of photoactive electrodes for a high-
efficiency DSC. FTO glass is used as current collector (Nippon Sheet Glass, solar
4-mm thickness), which can be cut into the desired size by using a diamond glasscut-
ter and a glass breaker (Fig. 8.4). To prepare the DSC working electrodes, the FTO
glass is first cleaned in a detergent solution using an ultrasonic bath for 15 minutes

Fig. 8.3 Fabrication scheme of dye-sensitized-Ti02 electrodes.


How to make high-efficiency dye-sensitized solar cells 255

Fig. 8.4 A diamond glass cutter (a) and a glass breaker (b). With these tools, 4-min-thickness
glass substrate can be cut into the test pieces (c).

(Fig. 8.5), and then rinsed with tap water, pure water and ethanol (Fig. 8.6). After
treatment in an UV-0 3 system for 18 minutes (Fig. 8.7), the FTO glass plates are
immersed into a 40 mM aqueous TiCl4 solution at 70 °C for 30 minutes (Fig. 8.8)
and washed with pure water and ethanol and dried (Fig. 8.6 (b) and (c)). A layer of
nanocrystalline-Ti02 paste (anatase, d = 20 nm) was coated on the FTO glass plates
by screen printing (Fig. 8.9).
The details of screen-print procedure are illustrated in Figure 8.10. The angle
of the coating rubber blade has to be kept as shown in this figure, otherwise the Ti0 2
paste leaks out the side of the mask, and layers like those in Figure 8.11 are pro­
duced. These would not be suitable for the precise measurement system [8.10] and
the fabrication of high-efficiency DSCs. The screen-printed substrates are kept in a
clean box saturated with ethanol for approximately six minutes, to reduce the sur­
face irregularity of the coated paste (levelling) (Fig. 8.12); these are then dried for
6 minutes on a hot plate at 125 °C. However, the exact leveling time is controlled
by the operator, because the leveling speed depends on the viscosity of each paste.
This screen-printing procedure with the paste (coating, storing and drying) is repeated
to obtain an appropriate thickness for the working electrode. After drying the paste
at 125 °C, two layers of the Ti0 2 paste are deposited by screen printing to make a
256 Dye-Sensitized Solar Cells

Fig. 8.5 View of an ultrasonic-bath washing. A plastic box with detergent solution and FTO
substrate (a) was put in a sonicator (b).

Fig. 8.6 Washing procedure with tap water (a), pure water and ethanol (b) and dry (c).

Fig. 8.7 UV-0 3 cleasing: the survey (a) and the sample stage and FTO substrates (b).
How to make high-efficiency dye-sensitized solar cells 257

Fig. 8.8 TiCl4 treatment by chemical-bath deposition: Dipping FTO substrate into 40 mM
TiCl4 (aq.) bath in a plastic box (a) and heating the bath in an oven at 70 °C.

Fig. 8.9 Procedure of screen printing: Putting a Ti02 paste on a screen mask (a); sliding a rub­
ber blade for the screen printing (b); and the end of the printing (c).

light-scattering-Ti02 film consisting of 400-nm-sized anatase particles, resulting in


a total thickness of 4-5 μΐη. The electrodes coated with the Ti0 2 pastes are gradually
heated under an air flow according to the following schedule: 325 °C for 5 minutes, at
375 °C for 5 min, at 450 °C for 15 min, and 500 °C for 15 min.
The sintered Ti0 2 film was treated again with 40 mM TiCl4 solution as described
above (Fig. 8.8), rinsed with pure water and ethanol (Fig. 8.6 (b) and (c)) and sin­
tered again at 500 °C for 30 minutes. After cooling to 80 °C, the Ti0 2 electrode is
immersed into a 0.5 mM N-719 dye solution in a mixture of acetonitrile and tert-butyl
alcohol (volume ratio: 1:1) and kept at room temperature for 20-24 hours to complete
the sensitizer uptake.
258 Dye-Sensitized Solar Cells

Fig. 8.10 Screen-printing procedure:filling-ina Ti02 paste (a)-(c) and printing the paste on a
substrate (d)-(f). The angles of the rubber blade are 85°-88° and 70°-80° from the horizontal at
thefilling-inand the printing, respectively

8.2.4 Counter-Pt electrodes


To prepare the counter electrode, a hole is drilled in the FTO glass (LOF Industries,
TEC 15 Ω/D, 2.2 mm thickness) by sand blasting. The perforated sheet is washed
with H 2 0 as well as with a 0.1 M HC1 solution in ethanol and cleaned by ultrasound
in an acetone bath for 10 minutes. After removing residual organic contaminants by
heating in air for 15 min at 400 °C, the Pt catalyst is deposited on the FTO glass by
doctorblade coating with a drop of H2PtCl6 solution (2 mg Pt in 1 ml ethanol) and
repeating the heat treatment at 400 °C for 15 minutes.

8.2.5 DSC assembling


The dye-covered Ti0 2 electrode and Pt-counter electrode were assembled into a
sandwich-type cell (Fig. 8.13) and sealed with a hot-melt gasket of 25-μιη thick­
ness made of the ionomer Surlyn 1702 (Dupont) on a heating stage. Then, the
hole in the back of the counter electrode is covered with a hot-melt ionomer film
(Bynel 4164, 35 μιη thickness, Du-Pont) by using a hot soldering iron covered by a
How to make high-efficiency dye-sensitized solar cells 259

Fig. 8.11 Photographs of screen-printed Ti02filmswith different angles of rubber blade action:
(a) proper angle; and (b) lower angle of about 45°.

Fig. 8.12 (a) Illustration of setup of Ti02-paste leveling in a plastic box with a sheet of ethanol-
soaked tissue paper for 3 min; surface profiles of Ti02 layers without (b) and with (c) leveling.

fluorine-polymer film. A hole was made in the hot-melt ionomer film with a needle.
A drop of the electrolyte, a solution 0.60 M butylmethylimidazolium iodide [8.11],
0.03 M I2, 0.10 M guanidinium thiocyanate and 0.50 M 4-^ri-butylpyridine in the
mixture of acetonitrile and valeronitrile (volume ratio: 85:15) is placed onto the hole.
The cell is put into a small vacuum chamber to remove inside air in a few seconds.
Exposing it again to ambient pressure causes the electrolyte to be driven into the cell
by the process of vacuum back-filling. Finally, the hole is covered by an additional
hot-melt ionomer film (Bynel 4164, 35-μιη thickness, Du-Pont) and a cover glass
(0.1-mm thickness), and sealed by using a hot soldering iron. In order to have good
260 Dye-Sensitized Solar Cells

Fig. 8.13 Configuration of the DSC.

electrical contact for the connections to the photovoltaic measurement setup, the
edge of the FTO outside of the cell is scraped slightly with sandpaper or a file. A
solder (Cerasolza, Asahi Glass) was applied on each side of the FTO electrodes with
an ultrasonic-soldering system.

8.2.6 Measurements
Photovoltaic measurements employed an AM 1.5 solar simulator (lOOmWcnr 2 ).
The power of the simulated light was calibrated by using a reference Si photodiode
equipped with an IR-cutoff filter (KG-3, Schott) in order to reduce the mismatch in
the region of 350-750 nm between the simulated light and AM 1.5 to less than 2 %
[8.12, 8.13]. I-V curves were obtained by applying an external bias to the cell and
measuring the generated photocurrent with a Keithley model 2400 digital-source
meter.

8.3 RESULTS AND DISCUSSION

8.3.1 TiCl4 treatments


Electrodes are treated twice with TiCl4, once before and once after the porous-Ti02
screen printing (Fig. 8.3). The first TiCl4 treatment has two effects: to enhance the
bonding strength between the FTO substrate and the porous-Ti02 layer, and to block
How to make high-efficiency dye-sensitized solar cells 261

charge recombination between electrons in the FTO with holes in Γ/Ι3 redox couple.
The second TiCl 4 treatment has the effect of enhancing the surface roughness factor
for dye adsorption, resulting in high photocurrent [8.6]. Figure 8.14 shows the dark
current-voltage characteristics of the two kinds of flat electrodes with and without
TiCl 4 treatment. The onset of the dark current of the FTO electrode occurs at low
forward bias. The TiCl 4 treatment suppresses the dark current, shifting it by several
hundred millivolts. As a result, it can work at the lower voltage region. This indicates
that the reduction of triiodide at the exposed part of FTO is responsible for the high
dark current.
Table 8.15 shows the influence of TiCl 4 treatment on the electrode charac­
teristics [8.6]. Although the specific surface area was decreased 7.9 % by the TiCl 4
treatment, the T i 0 2 weight increased by 28.1 %, resulting in a 19.0 % increase of
the T i 0 2 roughness factor. Consequently, the absorbance increased 15.7 %, thanks

Fig. 8.14 Dark I-V curves of flat electrodes without Ru-dye: FTO (dotted line) and FTO/Ti02
(solid line). Flat Ti0 2 was made by double TiCl4 treatment on FTO. The area of electrodes is
0.36 cm2.

Table 8.15 Characteristics of nanocrystalline Ti0 2 layers with and without TiCl4 treatment.
Each number is calculated to "par 1 μιη", and is the average of three samples, except for the
BET measurements that provide the specific surface area and average pore size [8.6].

Electrodes nano-TiOo TiCl4-treated nano-Ti02

Average pore diameter (nm) 20.2 18.3


Specific surface area (m2 g_1) 86.0 79.7
Ti0 2 weighta (mg cm -2 μητ 1 ) 0.135 + 0.003 0.173 + 0.003
Roughness factorb (μητ 1 ) 116 + 3 138 + 2
Absorbance at 540c (nm/μητ1) 0.159 + 0.05 0.184 + 0.06
a
The weight-measurement sample area was 16 cm2 with 15 μιη thickness.
b
The roughness factor was obtained by multiplying specific surface area and Ti02 weight.
c
Absorbance measurements were performed with N719-adsorbed nanocrystalline Ti02 layer at 540 nm.
The back ground was the same Ti02 electrode after removal of Ν719 by 0.1 Mteri-buthylammonium
hydroxide in acetonitlile. The pores in nanocrystalline Ti02 layers werefilledwith butoxyacetonitrile to
decrease the light scattering effect.
262 Dye-Sensitized Solar Cells

to the TiCl 4 treatment. The increase of diameter (1.9 nm) after the TiCl 4 treatment
suggests the generation of an additional T i 0 2 layer (l-nm thickness) on the surface of
nanocrystalline T i 0 2 in the porous layer (Fig. 8.16). With the increases of roughness
factor and dye absorbance, the photocurrent and the conversion efficiency increased
by 9.6 % and 8.0 %, respectively. Therefore, DSCs with conversion efficiency exceed­
ing 10 % can be obtained (Fig. 8.17).

8.3.2 Effect of the light-scattering Ti0 2 layer


In order to enhance the photocurrent of DSCs, a photon-trapping system has been
applied to porous T i 0 2 electrodes by the use of transparent and light-scattering lay­
ers, the so-called double-layer system [8.6, 8.7]. Figure 8.17 shows the effect of the
double-layer system. Without the light scattering layer, the photovoltaic character­
istics were, / s c = 15.6 mA cm- 2 , Voc = 791 mV, FF = 0.740 and η = 9.12 %. When
the light-scattering layer is added, the photovoltaic characteristics were enhanced
to /sc = 18.2 mA cm"2, V o c = 789 mV, FF = 0.704 and η = 10.1 %. Earlier it was

Fig. 8.16 Effect of TiCl4 treatment: an additional Ti0 2 layer (l-nm thickness) was coated on
the surface of nanocrystalline Ti0 2 in the porous film.

Fig. 8.17 I-V curves showing the different Ti0 2 electrodes prepared according to a standard
procedure (Fig. 8.3) and then without TiCl4 treatment and without the light-scattering layer.
The thickness of the transparent and light-scattering layers were 14 μπι and 5 μιη, respectively.
Photovoltaic characteristics: standard Ti0 2 electrode, / S c = 18.2 mA cm-2, Voc = 789 mV,
FF = 0.704 and η= 10.1 %; without TiCl 4 ,/ S c = 16.6 mA cm"2, VOC = 778 mV, FF = 0.731 and
77 = 9.40 %; without the light-scattering layer, / s c = 15.6 mA cm"2, V0c = 791 mV, FF = 0.740
and Ύ] = 9.12%.
How to make high-efficiency dye-sensitized solar cells 263

reported that the light-scattering layer was important not only for the photon-trapping
system, but also for the photovoltaic generation itself: the DSC with the dye-sensi­
tized light-scattering-Ti02 layer (without the transparent nanocrystalline-Ti02 layer)
is characterized by a 5 % conversion efficiency [8.14].

8.3.3 Thickness of the nanocrystalline Ti0 2 layer


In order to optimize the photovoltaic performances of DSCs, it is important to vary the
thickness of the nanocrystalline-Ti02, because the locus of photoconversion is at the
surface of the dye-covered Ti0 2 . The surface area can be calculated from the porous
layer thickness. The measurement of the exact thickness would be difficult after produc­
ing the double-layer electrode (composed of the transparent-nanocrystalline layer and
the light-scattering-submicrocrystalline Ti0 2 layer), and so we measure the nanocrystal-
line-Ti02 layer before sintering by use of a surface profiler. In order to project the result­
ing thickness of the sintered nanocrystalline-Ti02 layer in the double-layer electrode,
a calibration curve relating the thicknesses before and after sintering of the nanocrys-
talline-Ti02 layer was used (Fig. 8.18). The shrink ratio of thicknesses from "before
sintering" to "after sintering" was 0.942 for the layers with 20 nm-Ti02 particles. Using
the calibration line of Figure 8.18, the relationship of thickness of nanocrystalline-Ti02
layer and the conversion efficiency of DSC was obtained (Fig. 8.19). It is confirmed that
the 12-14 μιη thickness is the optimum for obtaining high efficiency DSCs.

8.3.4 Anti-reflecting film


Since glass substrates reflect 8-10 % of incident light, an anti-reflecting film is neces­
sary to enhance the photovoltaic performances of DSCs. The light-reflecting losses
are eliminated by using a self-adhesive fluorinated polymer film (ARCTOP, ASAHI
GLASS) that serves at the same time as a 380-nm UV cut-off filter (Fig. 8.20). Masks
made of black plastic tape were attached on the ARCTOP filter to reduce scattered
light. Figure 8.21 shows the difference of incident photon-to-electron conversion

Fig. 8.18 Calibration line for projecting thicknesses of sintered transparent nanocrystalline-
Ti02 electrode by measuring the thickness before sintering.
264 Dye-Sensitized Solar Cells

Fig. 8.19 Relationship of thickness of transparent nanocrystalline-Ti02 layer and the conver­
sion efficiency of DSC with anti-reflecting films. Each point is the average of four cells.

Fig. 8.20 Tranmittance measurements of a glass substrate with and without an anti-reflection
film (ARCTOP).

efficiency (IPCE) upon application of the anti-reflecting film, using double-layer


electrodes (14 μιτι and 5 μιη thicknesses of transparent and light-scattering T i 0 2 lay­
ers). The anti-reflecting film enhance the IPCE from 87 % to 94 %, resulting in the
improvement of the conversion efficiency by 5 %.

8.3.5 Reproducibility of DSC photovoltaics


DSCs with over 10 % efficiency are fabricated after optimization of thickness of
nanocrystalline-Ti0 2 layer, TiCl 4 treatments and application of an anti-reflecting film.
Figure 8.22 shows the statistics of photo-to-electricity conversion efficiency of DSCs
How to make high-efficiency dye-sensitized solar cells 265

Fig. 8.21 Effect of anti-reflecting film (ARCTOP) on the IPCE of DSC.

Fig. 8.22 Efficiencies of 12 DSCs made on the 15th and 16th March, 2005.

made at the same time. Out of the 12 devices, all showed 10 % efficiency or greater.
Considering the measurement error with a solar simulator, it is concluded that repro­
ducible values of 10.2 ± 0.2 % is attainable by these techniques.

8.4 CONCLUSION

We describe the important points for the reproducible fabrication of high-efficiency


DSCs, particularly in terms of nanocrystalline-Ti02 layer thickness, TiCl4 treatment
and the application of an anti-reflection layer. Other important points are the selec­
tion of FTO glass for the working electrode (NSG, Solar-4 mm) and counter elec­
trode (LOF, TEC 15-2.2 mm), and the selection of the ultrasonic solder. Moreover,
the fabrication of the counter-Pt electrode, the purification of Ru dye (N719) and the
composition of electrolyte all contribute to the high performance of the devices. The
combination of all of these state-of-the-art practices is responsible for the production
of reproducible DSCs with greater than 10 % efficiency.
266 Dye-Sensitized Solar Cells

8.5 ACKNOWLEDGEMENTS

This work was supported by a grant from the Swiss Federal Energy Office (OFEN).

8.6 REFERENCES

[8.1] O'Regan, B.; Grätzel, M. Nature, 1991, 335, 737.


[8.2] Grätzel, M. Nature, 2001, 474, 338.
[8.3] Hagfeldt, A.; Grätzel, M. Ace Chem. Res., 2000, 33, 269.
[8.4] Bach, U.; Lupo, D.; Comte, P.; Moser, J. E.; Weissörtel, F.; Salbeck, J.; Spreitzert, H.; Grätzel, M.
Nature, 1998, 395, 544.
[8.5] Smestad, G.; Solar Energy Mater. Solar Cells, 1994, 32, 259.
[8.6] Ito, S.; Liska, P.; Charvet, R.; Comte, P.; Péchy, P.; Nazeeruddin, Md. K.; Zakeeruddin, S. M.;
Grätzel, M. Chem. Commun., 2005, 4351.
[8.7] Wang, P.; Zakeeruddin, S. M.; Comte, P.; Charvet, R.; Humphry-Baker, R.; Grätzel, M. /. Phys.
Chem. B, 2003, 707, 14336.
[8.8] Nazeeruddin, Md. K., Zakeeruddin, S. M.; Humphry-Baker, R.; Jirousek, M.; Liska, P.;
Vlachopoulos, N.; Shklover, V.; Fischer, Ch.-H.; Grätzel, M. Inorg. Chem., 1999, 38, 6298.
[8.9] Nazeeruddin, Md. K.; De Angelis, F ; Fantacci, S.; Selloni, A.; Viscardi, G.; Liska, P.; Ito, S.;
Takeru, B.; Gratzel, M. /. Am. Chem. Soc, 2005, 727, 16835.
[8.10] Ito, S.; Nazeeruddin, Md. K.; Liska, P.; Comte, P.; Charvet, R.; Péchy, P.; Jirousek, M.; Kay, A.;
Zakeeruddin, S. M.; Grätzel, M. Progress in Photovoltaics., 2006, 74, 589.
[8.11] Bonhôte, P.; Dias, A. P.; Armand, M.; Papageorgiou, N.; Kalyanasundaram, K.; Grätzel, M. Inorg.
Chem., 1996,55,1168.
[8.12] Nazeeruddin, Md. K.; Péchy, P.; Renouard, T.; Zakeeruddin, S. M.; Humphry-Baker, R.; Comte, P.;
Liska, P.; Cevey, L.; Costa, E.; Shklover, V.; Spiccia, L.; Deacon, G. B.; Bignozzi, C. A.; Grätzel,
M. /. Am. Chem. Soc, 2001,123, 1613.
[8.13] Ito, S.; Matsui, H.; Okada, K.; Kusano, S.; Kitamura, T.; Wada, Y.; Yanagida, S. Solar Energy
Mater. Solar Cells, 2004, 82, All.
[8.14] Zhang, Z.; Ito, S.; O'Regan, B.; Kunag, D.; Zakeeruddin, S. M.; Liska, P.; Charvet, R.; Comte,
P.; Nazeeruddin, Md. K.; Péchy, P.; Humphry-Baker, R.; Koyanagi, T.; Mizuno, T.; Grätzel, M.
Zeitschrift für Physikalische Chemie, submitted 2009.
CHAPTER 9

SCALE-UP AND PRODUCT-DEVELOPMENT


STUDIES OF DYE-SENSITIZED SOLAR CELLS
IN ASIA AND EUROPE
K. Kalyanasundararn, Seigo Ito, Shozo Yanagida and Satoshi Uchida

9.1 INTRODUCTION

The market for photovoltaic solar cells can be divided broadly into two categories with
the largest share for large-area installations intended for power generation. Electric
power can also be generated and used locally (off-grid sources to supply houses or run
irrigation pumps in remote areas), or it can be used to feed to a national power grid. A
smaller but growing share is based on small-area solar cells (sub-modules, typically
< 100 cm2) for use in portable electronics such as laptop computers, solar bathroom
balances, mobile-phone chargers, solar garden lamps and similar uses. Development
work has been taking place in both areas. Table 9.1 lists some of the growing types of
applications of dye-sensitized solar cells (abbreviated as DSC) under development in
different laboratories.
An important point to note in photovoltaic cell studies is the size of the solar
cells being examined. Nearly all fundamental and optimization studies in laboratories
use small-area cells with illumination area of < 1 cm2. The first stage of scaling up is
to increase the active area by orders of magnitude in modules. Here the illumination
area can be anywhere in the range of 25-100 cm2. Much larger-area installations for
power generation use solar panels (1 m2 or larger). Panels are formed by combining
several modules. Solar cells in general are sold in the form of sub-modules, modules
and panels. The term solar array refers to using several large area solar panels in
large-area field installations.
In this chapter, we highlight various product development studies on DSCs
in different industrial laboratories of Asia and Europe. This chapter is complemen­
tary to the earlier chapter on packaging, scaling and commercialization of DSCs by
Desilvestro et al. (Chap. 7 of this volume).
268 Dye-Sensitized Solar Cells

Table 9.1 Some of the application areas of DSC under development in various countries.

Type of application Companies

1. large-area solar panels for terrestrial power Sharp, Sony, Nanomax, IPP, FIS
generation (off grid, feed to local power grid)
2. a) large-area solar panels as part of Color-Sol project of Germany
building-integrated photovoltaics (BIPV) Dyesol Italian BIPV Company
b) large area panels to power electric cars Taiyo Yuden
c) panels to power military drones US Air Force
3. flexible light-weight portable power pack Peccell, G24Innovation
4. flexible power plastics Dyesol
5. power generation for remote area needs 3G Solar of Israel
6. street lights and road signs S JC - Shimane Inst. of Industrial Tech.
7. consumer household electronics Sony's self-powering lamp shades
J Touch solar-powered clocks

9.2 SCALING UP OF LABORATORY CELLS TO MODULES


AND PANELS

Development of modules is an important step in the advancement of any solar-cell


device destined for practical applications. For all photovoltaic devices, the highest
solar conversion efficiencies are obtained for small-area lab-test solar cells. When
the illumination (charge generation and collection) area is made larger, invariably
there is decrease (by a few percent) in the overall conversion efficiency per square
centimeter. This is due to loss of some of the photogenerated charge carriers via
recombination and/or trap processes. Losses in efficiency from lab-size cells to mod­
ules (several percent) have been noted in nearly all semiconductor-based PV systems
(single crystal-Si, amorphous Si, CdTe, CIGS). Reducing substantially this scale-up
loss is essential. For DSCs, some of the attempts in this regard involve optimization
of the materials preparation; of the mode of deposition of the active oxide layers; of
the counter-electrode design, and of the ways of inter-connecting small-area cells
to modules. Efficiency of the solar cell has to be above a certain threshold before
they are scaled up for pilot-level testing by major players of the photovoltaic (PV)
industry. Based on cost-estimates, solar conversion efficiency of 10-12 % is consid­
ered the absolute minimum in order for DSCs to be competitive with alternate PV
technologies.
Solar cells for portable electronics work invariably at ambient temperature
and at a light flux less than the full solar level. Requirements for thermal stability is
reduced for this type of application. Large-area solar panels employed out of doors
(terrestrial or outer-space) have to sustain more stringent operating weather condi­
tions. In tropical or high-altitude installations, panel temperatures during the year
may from -10-60 °C with the solar cells exposed to high humidity (> 75 %). To be
cost effective, they have to work reliably for long periods (product lifetime 15 years
or longer). Depending on the location, the solar panel may be exposed to full solar
radiation, or it may have to function under cloudy sky conditions for most of the year.
Scale-up and product-development studies of dye-sensitized solar cells 269

Constant improvements in the design and performance of portable electronic devices,


thanks to technological advances, have resulted in shorter periods for the guaranteed
cost-effective performance for portable electronics (10 years or even less). For dye-
sensitized solar cells, a product performance guarantee corresponding to a 20-year
lifetime requires several million turnover cycles of the key components.
The most important parameter for cost reduction is the overall sunlight-to-elec­
trical conversion efficiency. During the past decade, there has been growing sensi­
tivity in the general public of industrialized economies about energy consumption.
Increasingly, photovoltaic cells are integrated into building architecture, an area
known as building-integrated photovoltaics (BIPV). As for large-area installations
designed for power generation, higher efficiencies will reduce the area of the solar
panel required. For certification on reliable performance, solar-cell modules have to
be tested according to certain specifications. Some of the standards that apply to the
testing of thin-film solar panels and photovoltaic modules used in DSC development
include: IEC 61215 (terrestrial crystalline silicon photovoltaic (PV) modules, type
suitability and type approval); IEC 61646 (terrestrial thin-film photovoltaic PV mod­
ules, type suitability and type approval, JIS C-8938 standard Japanese counterpart)
and ASTM E 1171 (standard test method for photovoltaic modules in cyclic tempera­
ture and humidity environments).
The following thermal stability tests are part of these standards:

• thermal-cycling test: cycling between 85 °C and -40 °C at 100 °C/hour maxi­


mum (to assess the module's ability to withstand exposure to several environ­
mental conditions during transportation and/or storage);

• temperature-humidity cycling test: cycling between +85 ± 2 °C and -40 ± 3 °C


at 85 ± 5 % RH, at 100 °C/hour maximum (to determine the deterioration level
for use and/or storage in short time under conditions of temperature change in
high relative humidity);

• light soaking: exposure of the solar cell to continuous illumination of solar


radiation for 1000 hours.

Toyota/Aisin Seiki, Sharp, Shimane Institute of Technology and Fujikura are


some of the DSC module developers who have already optimized their cell design
and packaging to pass performance tests involving international and national stand­
ards, such as IEC 61646 "thin-film terrestrial photovoltaic (PV) modules" or Japan
Industrial Standard C-8938. This underscores how DSC technology is rapidly evolv­
ing toward full commercial viability.
Since the seminal report of O'Regan and Grätzel in 1991 in the journal Nature
[9.1], there have been numerous exploratory studies using a wide variety of dyes,
redox mediators, electrolytes, oxide substrates, counter-electrodes and even compo­
nent assembly modes. These studies have led to a systematic increase in the over­
all light conversion efficiency, currently around 12 % for small area lab-cells and
8 % for modules. The stability of dye-sensitized solar cells over extended photolysis
periods has improved sufficiently to pass the international standards for thin-film PV
modules. Encouraged by these developments, several laboratories have undertaken
270 Dye-Sensitized Solar Cells

development of larger-area modules for practical applications, mainly over the past
decade. Over one hundred industrial laboratories worldwide, big and small, are now
engaged in the development of DSC-based photovoltaic power generation systems.
Substantial amounts of money are being invested in new start-ups by venture invest­
ment agencies. In this chapter, we review the state of the art in some of the leading
industrial laboratories, most of them located in Europe and Asia. The review intends
to illustrate the challenges faced and the novel approaches taken by different indus­
tries to address them.

Module design considerations


It was mentioned earlier that fundamental studies to identify key elements of solar cell
performance are done in small-area cells with a surface less than one square centim­
eter. Key parameters for optimization are incident monochromatic power conversion
efficiency (IPCE), maximum photocurrent (/sc) and photo voltage (Voc) at optimum
power point and the fill factor (ff). Overall solar-to-electrical conversion efficiency is
based on these fundamental parameters. Solar cells size and conductivity of the sub­
strates influence significantly the internal resistance of solar cells and consequently
the fill factor and the conversion efficiency of the DSCs. Even for cases where the
solar conversion efficiency is fairly high (> 8 %), simple scaling of area to modules
of 10 x 10 cm deposited on TCO glass without any collector electrode yield power
conversion of 1-2 % only. Printing silver finger as internal current collector electrode
(reducing the charge carrier collection area to cm2 or less) and the manner in which the
individual cell elements are connected thus are very important. For DSCs, three dif­
ferent types of connecting small area cells have been studied: (i) the "series-Z" design
as used in the early studies of modules by Sustainable Technologies International
STI/Dyesol of Australia; (ii) the parallel or "masterplate" design as used by the Dutch
Energy research laboratory ECN; and (iii) the monolithic design as proposed initially
by Andreas Kay of EPFL and developed by Aisin Seiki and others particularly in
Japan. Depending on the module design and inter-connect of constituent cells, effec­
tive area for power generation (with respect to outer geometric area of the module)
can be anywhere between 70-90 %.
Modules with series interconnections has provided the best route to industrial
production. The Z-series inter-connect design (introduced by STI), can be used for
glass/metal, plastic/metal, glass/plastic and plastic/plastic substrates, and is com­
prised of two opposing electrodes with the connection between cells consisting of
a conducting medium. The advantage of this design is its high-voltage output with
relatively small interconnect-resistance losses, and its facility for pre- and post-treat­
ment of the working electrode. The disadvantage is the risk of a lower fill factor,
which results from the series resistance of the interconnect electrode. STI selected
this design after the invention of an interconnect design with low resistance. The
working electrode and counter-electrode can be optimized separately, and there is
no requirement to mask the counter-electrode or pre-seal the module when apply­
ing the dye. Consequently, the dye uptake can be more carefully controlled on the
basis of manufacturing cost analysis and reproducibility. Research work at ECN led
by Jan Kroon developed the 'masterplate' concept, and this has been used in several
Scale-up and product-development studies of dye-sensitized solar cells 271

European projects. The monolithic module (or Kay cell named after the inventor Dr.
Andreas Kay of EPFL) is now becoming the major design model for volume produc­
tion and for relatively small cells.

9.3 DSC DEVELOPMENT STUDIES IN VARIOUS EUROPEAN


LABORATORIES

A large number of industrial laboratories in Europe have been studying both the
fundamentals and scaling up of DSCs for nearly two decades. Some of the lead­
ing laboratories deserve mention: the Institute of Applied Photovoltaics (INAP,
Gelsenkirchen, Germany); the Energy Research Center of Netherlands (ECN, Petten,
Netherlands); IMRA-Europe (Sophia Antipolis, France); Solaronix (Aubonne,
Switzerland); Fraunhofer Institute for Solar Energy (ISE, Stuttgart, Germany); Sony-
Europe Research Center (Stuttgart, Germany); G24Innovation (Cardiff, Wales, UK);
Greatcell S.A. (now owned by DYESOL/STI of Australia); and Solterra Fotovoltaico
S.A. (Chiasso, Switzerland and Dyesol Italia). Solterra was an industrial partner of the
laboratory of Professor M. Grätzel until 2001, when they shifted their plans to focus
on production of Si-solar cells. IMRA is the major research partner of Toyota and its
research wing Aisin Seiki. Developments from Aisin Seiki are discussed later on in
this chapter, along with the work of Sony.

9.3.1 Energy Research Centre of the Netherlands (ECN)


The Energy Research Centre of the Netherlands (ECN) was one of the first European
Laboratories to take development work of DSC technology to sub-modules, with
work beginning in 1995 [9.2-9.10]. They have developed the so-called masterplate
design for DSC modules. In their collaborative efforts with several European partners,
they have prepared DSCs in various colors. Figure 9.2 shows one such masterplate
DSC module and a multicolor panel. ECN Researchers reported results in 2003 on
their semi-automated system for reproducible manufacturing of DSCs on sizes up to
100 cm2. Manufacture of two types of glass-glass DSCs were examined: small-area
cells (< 5 cm2) with a conversion efficiency of 5.9 %. These cells had an active area/
total area ratio of 0.68, translating into a device efficiency of 4.3 %. Batch production
of 27 cells per day was successful (26 of 27 cells with satisfactory yield and 22 out
of 27 with good reproducibility). Average I-Vparameters for the 26 cells (with active
area of 68 cm2) was found to be: Voc = 0.68V, 7SC = 10 mA/cm2, ff= 0.62, and the effi­
ciency was measured to be 4.3 % (active area).
During the two-year period 2002 to 2004, a consortium of four European uni­
versities (EPFL, Imperial College, Cracow University, Materials Research Center of
Freiburg), three research institutes (ECN Solar Energy of Netherlands, Fraunhofer
ISE of Germany, IVF Industrial Research and Development Corporation) and one
industrial partner (Greatcell Solar S.A.) cooperated under a European project called
NANOMAX. The goal of the NANOMAX was to test new strategies for DSC cell
design, cell materials and fabrication protocols with the aim to increase the efficiency
272 Dye-Sensitized Solar Cells

Fig. 9.2 Masterplate design of DSC of ECN consisting offivesingle cells, 2.5 cm2 each, on a
7.5 x 10 cm2 surface sandwiched between two FTO coated glass substrates (left); DSCs in dif­
ferent colors (right). Courtesy of Dr. Winfried Hoffmann, RWE Schott Solar, Germany.

to above 12 % under standard test conditions (AM 1.5, 1000 W/m2) with good long-
term stability. In addition, cost analyses were made to demonstrate the potential of
DSCs as a low-cost thin-film PV technology.
The combined research efforts have led to the following technical
achievements:

New ruthenium-containing sensitizing dyes with enhanced optical absorption


in the visible part of the spectrum have become available and have been suc­
cessfully applied in DSCs;

Protocols for making metal-oxide blocking layers on Ti0 2 result in the retarda­
tion of recombination dynamics and improvement of the photovoltage of the
device. New concepts, such as the TCO-less design, have been developed and
introduced. New scatter phenomena in Ti0 2 films have been discovered.

A maximum power conversion efficiency under full sunlight of 11 % for areas


of approximately 1 cm2 has been achieved.

A cell with efficiency exceeding 8 % and that retains over 98 % of its initial
performance after 1000 hours of accelerated testing, and under thermal stress
at 80 °C in the dark, has been demonstrated. Negligible device degradation was
observed for 1000-h visible-light soaking at 60 °C. Long term stability at ele­
vated temperature has been achieved using hydrophobic Ru-dyes with pendant
alkyl chains, such as Z90 and K19.

Advanced techniques have been developed for in situ characterization of dye-


sensitized photovoltaic devices under operation, including transient optical
Scale-up and product-development studies of dye-sensitized solar cells 273

studies covering all key steps of interfacial charge separation and recombina­
tion dynamics. Also, transient photovoltage and photocurrent studies of trans­
port dynamics have been conducted. These experiments have been correlated
with I-V data in the dark and under illumination, as well as with a numerical
model of these I-V data based upon the non-ideal diode equation. These meth­
odologies have been employed to characterize fundamental loss factors in both
standard devices and in a range of innovative device concepts as developed in
this proposal. For standard devices the primary loss mechanism is confirmed to
be charge recombination of electrons in the metal oxide to the oxidized redox
couple.

• Modules of different designs (Z-type, current-collecting and monolithic) and


sizes (from 100 up to 900 cm2) have been demonstrated in the existing process­
ing baselines with maximum active area efficiencies of 5.5 % under conditions
of 1 sun and 6.5 % at 0.1 sun.

A European consortium financed under the Joule program (LOTS-DSC, JOR3-


CT98-0261) has confirmed the cell-photocurrent stability during 10,000 hours of light
soaking at 2.5 suns, corresponding to an approximate 56 million turnovers of the dye
without any significant degradation. A more difficult task has been to reach stability
under prolonged stress at higher temperatures, i.e., from 80-85 °C. Accelerated long-
term tests for a number of single cells containing different dye and electrolyte com­
binations have been performed at different temperatures up to 85 °C in the dark for
periods up to 1000 hours, as well as under simulator sunlight. The ageing experiments
that were conducted to determine the stability of the DSCs on an intrinsic, molecular
level. With the solar conversion efficiency of liquid-electrolyte dye cells reaching an
efficiency of 10 % and above in the laboratory (surfaces of about 1 cm2), ECN has set
its project goal to develop a solid-state dye-sensitized solar cell with an efficiency of
10 % and a stability of 10 years.

9.3.2 Fraunhofer Institute for Solar Energy Systems (Fraunhofer ISE)


Research efforts by Hinsch and coworkers at the Fraunhofer Institute for Solar Energy
Systems (ISE) has focused on the stability and scaling up of DSCs into large mod­
ules [9.8-9.10]. As the DSC technology progresses from laboratory-scale to large-
area applications, long-term stability is one major obstacle. Especially for large-area
DSC modules, stability is often a matter of providing hermetic sealing both between
cells and for the whole module. To these ends, glass frit has been studied as sealing
material. Glass frit is thermally, mechanically and chemically very stable and can be
applied with screen-printing techniques. The Fraunhofer ISE has succeeded in devel­
oping large 30 x 30 cm2 modules with internal interconnections (meander type) using
glass-frit-sealing technology. Figure 9.3 shows one such glass-frit-based DSC mod­
ule. The decorative design has been achieved by screen printing a structured light-
scattering layer (sun pattern) internally onto the photoelectrode layer. The optical
appearance results from back scattered light (a module efficiency of 4 to 5 % over the
total area can be realized on a short term with such concept. The thermal stability of
274 Dye-Sensitized Solar Cells

Fig. 9.3 Semi-transparent glass-frit-sealed dye-solar-cell modules (30 x 30 cm2 each) as devel­
oped at Fraunhofer ISE.

Table 9.4 Efficiency parameters for glass-frit-based DSCs of the Fraunhofer Institute.
Intensity he (mA) Voc (V) FF (%) 7J(%) -* module V *-v

Sulfur lamp 1 sun 945 4.55 56 4.6 45


Roof 992 W/m2 830 4.66 57 4.2 34

the glass-frit sealing and the integrated-series connections was verified in a thermal
cycling from 40 to 80 °C. The coloration process has been scaled up to 30 x 30 cm2 by
pumping the dye solution through the module using only two filling holes. By heating
the module to 70 °C, they were able to fill the module with electrolytes based on high
viscous ionic liquids.
Table 9.4 below provides results on the efficiency measurements on a typical
glass-frit-sealed, semi-transparent DSC module under artificial light and outdoor illu­
mination (Determined for an active module area of 520 cm2, the total module area
without frame being 672 cm2).
Semi-transparent glass-frit-sealed DSC modules have been characterized on a
roof-top test site at the Fraunhofer ISE (see Fig. 9.5). The measurements are compared
to data from a-Si and CIGS commercial modules recorded at the same site. The modules
were mounted at a 45° tilt angle facing south. I-V data were recorded every 15 minutes.
The modules are kept in the open-circuit condition between the measurements.
Hinsch et al. have reported on the results achieved in the frame of a European
network project on integrated materials development for dye solar cells (21st European
Photovoltaic Solar Energy Conference, 4-8 September 2006, Dresden, Germany).
Using a simple milling process, a Ti0 2 screen printable paste has been developed
Scale-up and product-development studies of dye-sensitized solar cells 275

Fig. 9.5 Semi-transparent, glass frit-sealed Dye solar modules mounted on the roof-top of
Fraunhofer Institute, Germany for outdoor tests.

from in large scale commercially available Ti0 2 particles. Efficiencies up to 7 %


have been achieved using this paste. Electrolytes have been prepared from various
non-volatile ionic liquids. Mixtures of imidazolium based ionic liquids with differ­
ent anions resulted in a higher tri-iodide diffusion constant. Temperature dependant
measurements showed, that already at a cell temperature of 40 °C pure molten iodide
salts electrolytes can be applied in electrode distances up to 40 /mi without diffusion
limitation of the current (15 mA/cm2).
Quasi solid-state electrolytes have been successfully tested by applying organic
(PVDF-HFP) and inorganic (Si02) physical gelators. Based on a polyole process,
screen printed transparent catalytic platinum layers with extremely low charge-trans­
fer resistance (0.25 Ohm cm-2) have been developed. A conductive catalytic layer was
created by coating Sn02:Sb nanoparticles with platinum via hydrogen reduction, with
the aim to minimize the electrode distance in monolithic DSCs. This was done with
an optimized graphite paste, characterized by a sheet resistance of 6 Ohm cm -2 after
a high temperature treatment at 650 °C in inert atmosphere. This high temperature
stability allows the application of recently developed glass-frit sealing technology to
monolithic DSCs as well.
In a 2008 report, Hinsch et al. reported results of a German network project con­
ducted with twelve partners from universities and research institutes on the material
development of dye solar cells [9.9]. The goal of the project was to further evaluate
the concept of monolithic DSCs with respect to upscaling and reproducibility on glass
substrates. Methods were developed to provide a manufacturing process for mono­
lithic DSC modules based entirely on screen printing. Using the experience gained
with the sealing of standard DSCs, module encapsulation was achieved in a fusing
step by the soldering of glass-frit layers. For use in monolithic DSCs, a platinum-free,
conductive counter-electrode layer, characterized by a charge-transfer resistance of
276 Dye-Sensitized Solar Cells

RCT < 1.5 Vcm2, was developed by firing a graphite/carbon black composite under
an inert atmosphere. Glass-frit-sealed monolithic test cells have been prepared using
this platinum-free material. A solar efficiency of 6 % on a 2.0 cm2 active cell area has
been achieved in this case. Various types of non-volatile imidazolium-based binary
ionic-liquid electrolytes have been synthesized and optimized with respect to diffu­
sion-limited currents and charge-transfer resistances in DSCs.
Figure 9.6 shows another glass-frit-sealing-based DSC prepared in this man­
ner. Measurements have been done on so-called master plates; each plate consisting
of five individual cells with an active area of 2.0 cm2 (5 x 0.4 cm2). As transparent
conductive glass, TEC-8 (LOF, approx. 8 V/square) has been used. The screen print­
able pastes applied for the monolithic cells were prepared in a manner similar to the
method described earlier.
The best results achieved so far with this technology have been with a screen-
print paste (7 % efficiency, glass-frit sealed), dispersed by mechanical means in a pearl
mill (Getzmann) from commercial Ti0 2 particles (Kemira ANX type N, Finland).
The screen-printed photoelectrode layer was prepared from commercial Ti0 2 pow­
der (Kemira ANX type N). K19 dye (EPFL-Lausanne) was used, and the electrolyte
consisted of 0.8 M propylmethylimidazolium iodide (PMII); 0.15 M iodine; 0.5 M
Af-methylbenzimidazole (NMBI); 0.1 M guanidinium thiocyanate (GuSCN); and with
methoxypropionitrile as the volatile solvent. Experiments provided the following rep­
resentative I-V data for the solar cell: V oc = 752 mV; 7Sc = 15.2 mA/cm2; ff= 0.62;
area = 2.15 cm2; and the efficiency η = 7.12 %. In addition, quasi-solid-state electro­
lytes were successfully tested by applying inorganic (Si02) physical gelators. For the
use in semi-transparent DSC modules, a polyol process was developed, providing
screen-printed, transparent catalytic-platinum layers with the extremely low charge-
transfer resistance of 0.25 Vcm2.

Dye solar modules for facade applications: recent results from project ColorSol
With the ever-increasing interest on energy conservation, there are now con­
certed efforts to include photovoltaic power-generating systems as part of building

Fig. 9.6 A glass-frit-sealed dye solar cell with integrated series connections of Fraunhofer
ISE.
Scale-up and product-development studies of dye-sensitized solar cells 277

architecture - a field known as building integrated photovoltaic^ (BIPV). There are


two key features of DSCs, namely their optically translucent ("see-through") nature
and the choice of many colors, that make the DSC the superior choice for use in
BIPV. The Fraunhofer Institute has been involved in a German Industries Consortium
project known as ColorSol. The project focuses in particular on the application of
DSCs in BIPV (facades, PV-glazing, etc.). Various design concepts, as well as sce­
narios for the application of the DSC technology in architecture and facade plan­
ning were examined, and the application potential was quantified in cooperation with
potential users [9.11].
Figure 9.7 shows a glass façade with integrated DSC modules as displayed
at the Fraunhofer Booth during the 2007 European Photovoltaic Conference.
Prototypes of the glass-facade elements (70 x 200 cm2) were developed, consist­
ing of several serial-interconnected DSC modules of dimension 30 x 30 cm2. The
results of module characteristics under various outdoor illumination conditions
and under partial shading, as well as visual impressions of the DSC façade, were
reported. Several glass-frit-sealed modules were connected in series and laminated
between architectural glass panels of size 200 x 60 cm2. A decorative design was
achieved by screen printing a structured back-scattering layer of white porous Zr0 2

Fig. 9.7 Photograph of DSC demonstration module with a glass façade, as presented at the
Fraunhofer booth during the 22nd European Photovoltaic Solar Energy Conference (Milan,
Italy 2007).
278 Dye-Sensitized Solar Cells

(disk pattern) internally onto the transparent Ti0 2 photoelectrode layer prior to the
coloration with dye.

9.3.3 G24 Innovation


G24 Innovations Limited (referred to here as G24i) located in Cardiff, Wales, UK is
a major European undertaking devoted exclusively to the development of portable
power packs based on DSC technology. The DSC fabrication plant covers 18,000
square feet, set on 9.3 hectares, and it was commissioned in 2007. The G24i factory
will be the first facility to produce solar cells while relying exclusively on renewable
energy, including solar, wind, geothermal and other green sources. An automated roll-
to-roll manufacturing process transforms a lightweight roll of metal foil into a 50-
kilogram 800-meter-long roll of G24i dye-sensitized thin film in less than three hours.
G24i initially announced a DSC module production of 25 MW capacity in 2007 in
Cardiff, with plans to extend this to 200 MW by the end of 2008.
The G24i DSC thin-film design is uniquely thin, extremely flexible and versa­
tile nano-enabled photovoltaic (solar) material that converts light energy into electri­
cal energy, even under low-light, indoor conditions [9.12-9.14]. Thus, even in the
absence of direct sunlight, electric energy can be generated in both a convenient and
efficient manner. Figure 9.8 shows one of G24i's lightweight, flexible solar modules.
The company has incorporated the DSC thin film as part of an integrated universal
portable power charger, called the Power Curve Universal Solar Charger, now com­
mercially available.

Power Curve universal solar charger


Designed and engineered for rugged use in various applications, the Power Curve
solar charger has a rugged, durable, waterproof casing that can be attached to any

Fig. 9.8 The G24i dye-sensitized thin film (left); and the Power Curve universal solar power
charger of G24i (right).
Scale-up and product-development studies of dye-sensitized solar cells 279

outdoor gear. This device works in various light conditions, be it direct or indirect
sunlight, as well as with artificial light. This unit has the versatility to power mobile
telephones, AA batteries, digital cameras, MP3 players, and a family of personal com­
munication devices. The Power Curve unit features a battery with high-cycle longev­
ity along with the ability to charge both from the sun into the internal battery or via a
USB charging unit directly from a computer.
The company is joining with African companies, such as Kenya's MasterIT, to
distribute portable solar chargers locally, intent on solving the longstanding and vex­
ing problem of providing affordable light and electricity to poverty-stricken people
across Africa. Looking to bring off-grid electrical power options to people in Kenya,
Nigeria, Rwanda, South Africa and a still growing range of African countries, G24i
in May was awarded the World Bank Group's 2008 "Lighting Africa Development
Marketplace" prize for its solar-powered LED light, which uses dye-sensitized thin-
film solar cells in concert with light emitting diodes (LED) produced by Dutch
lighting manufacturer Lemnis. The award-worth $200,000-recognizes products that
provide the most innovative, off-grid lighting solutions. G24i is working with Lemnis
to carry out further development work and set up large-scale distribution in Rwanda,
the "Lighting Africa" program's initial target country.
In October 2009, G24i became the first manufacturer of dye solar cells to ship a
commercial product. The company's flexible modules are bound for Hong Kong-based
bag manufacturer, Mascotte Industrial Associates (MIA) to be incorporated into a line
of consumer electronics-charging knapsacks and business travel bags. Mascotte's line
of backpacks are capable of producing 0.5 W of power under direct sunlight, are are
therefore being targeted for on-the-go recharging of portable gadgets. Current designs
would allow full charge of a mobile telephone in 4-5 hours under sunny conditions;
indoor charging, depending on lighting, would take closer to 12 hours. Figure 9.9
shows some of the Mascotte's line of backpacks incorporating G24i's solar module.
In another developments, G24i and BASF are developing ionic liquids to fur­
ther boost the conversion efficiencies and long-term stability of G24i's solar cell tech­
nology. Recently G24i signed an agreement with the China National Academy of
Nanotechnology & Engineering (CNANE) in Tianjin, together with the Changchun

Fig. 9.9 Mascotte's line of backpacks incorporating G24i's solar module.


Next Page

280 Dye-Sensitized Solar Cells

Institute of Applied Chemistry (CIAC), which is part of the China Academy of


Sciences, and the Nanotechnology Industrialization Base of China (NIBC). The three
national Chinese institutes have agreed to commit resources to industrializing DSCs
with the objective of making significant advances in materials, manufacturing and
scientific aspects of G24i's thin-film solar technology.

9.3.4 3GSolar, Israel


3GSolar (formerly Orionsolar) is a leading developer of solar energy in Israel, using DSC
photovoltaic technology towards applications for off-grid rural areas. 3GSolar focuses
on rural off-grid markets, such as solar powered lamps and power supplies for remote
irrigation pumps, mainly in developing countries. Feasibility prototypes being devel­
oped are 15 x 15 cm cells, including robust current-collector grids and sealing methods.
Efforts are underway towards theirfirst-generationmodule of 55 W, composed of 15 cm
cells, each of 125 x 63 cm size, with a 7-year lifetime and 7 % efficiency.
3GSolara has developed a robust and intrinsically corrosion-resistant current-
collecting grid that allows scale-up to large area single dye cells of increased stability
and reduced cell inactive area. Prototype glass-based dye cells (Fig. 9.10) have been
scaled up to a full commercial single cell size of 225 square centimeters (aperture)
and presently give 5.4 % efficiency under one sun. These cells show good thermal
stability for 1500 hours at 85 °C. The modules have been constructed based on an
8 x 4 array of cells in series, and a double module of 64 cells is on test at a roof
station in Jerusalem. A single 32-cell module under one sun illumination delivers a
current of 3 A.
3GSolar is steadily ramping up cell efficiency and durability by optimization
of titania, dye, electrolyte, sealants and other cell components, and are on schedule
towards commercial production of low cost cells and modules starting 2011, initially
for the off-grid market in developing countries. The manufacturing line costs 40 %
less than the cost of a silicon line per megawatt output. Furthermore, the 3GSolar line
will operate efficiently at the size corresponding to 8 megawatts, meaning that they

Fig. 9.10 The 3GSolar dye cell prototype (225 cm2).


CHAPTER 10

CHARACTERIZATION AND MODELING


OF DYE-SENSITIZED SOLAR CELLS: A
TOOLBOX APPROACH
Anders Hagfeldt and Laurence Peter

10.1 INTRODUCTION

We begin this chapter by asking - what do we want to know about dye-sensitized solar
cells (DSCs)[10.1, 10.2]? This question can be answered at two levels. At a practical
level, we want to know what factors determine the performance of a DSC so that strat­
egies for improvement and optimization can be formulated. At a deeper, more fun­
damental, level, we wish to understand the physical and chemical processes involved
in the functioning of the DSC. The second objective presents a significant challenge,
since the 'chemical' processes taking place in the DSC differ substantially from those
in conventional solid-state photovoltaic cells. In fact, however, the underlying physics
of solid-state PV cells and DSCs is remarkably similar [10.3, 10.4]. Ultimately, a link
needs to be established between fundamental understanding and practical application,
since the raison d'être of DSC research is the development of a viable competitive
photovoltaic technology.
At a practical level, the important DSC performance parameters under standard
solar illumination conditions (usually AM 1.5) are the open circuit voltage, Voc, the
short circuit current density, j s c , and the fill factor, FF. The product of these three fac­
tors determines the solar power conversion efficiency, 77, when the cell is operating at
the maximum power point. The ratio of the short circuit current density to the incident
solar photon flux depends on the product of three factors - the wavelength-dependent
light harvesting efficiency, r/ih, the electron injection efficiency, 77^, and the efficiency
for collection of injected electrons, η00\. In the absence of light scattering, the esti­
mation of the light harvesting efficiency is relatively straightforward since it can be
calculated from the absorption properties and loading of the sensitizer dye. However,
high-efficiency DSCs use light-scattering layers to enhance performance, and in this
324 Dye-Sensitized Solar Cells

case calculations of η^ become more difficult. Reliable values of the injection and
collection efficiencies are also hard to obtain. T7inj depends on factors such as the
match between the energy levels of the photo-excited dye and the conduction band of
the mesoporous oxide (usually Ti0 2 ), the degree of electronic coupling between the
excited state of the dye and the conduction band states, and finally the competition
with other decay processes involving the excited state of the dye. T7col is sensitive to the
key loss processes - back electron transfer to I3 as well as to oxidized dye species. The
fill factor is probably the most poorly understood DSC parameter. It can be affected
by the series resistance and by sluggish electron transfer kinetics at the cathode, but it
can also reflect intrinsic non-ideality in the behavior of the DSC.
It is clear that a range of complementary experimental methods is required to
achieve a sufficiently detailed understanding of the DSC to allow the development of
realistic theoretical models that can be tested and refined for application in diagnostics
and optimization. This chapter describes a 'toolbox' approach [10.5] to achieving this
objective. The philosophy behind the toolbox approach involves recognizing that con­
sistent and comprehensive data can only be obtained if a wide range of experimental
methods is used to characterize to the same set of DSCs, since it is difficult to exactly
reproduce the preparative methods used in various laboratories.
The objective of the next section is to give a succinct summary of the cur­
rent understanding regarding the behavior of electrons in the mesoporous oxide in
order to provide the necessary background for the toolbox techniques, which are then
described in the subsequent parts of this chapter.

10.2 THEORETICAL BACKGROUND

10.2.1 Interfacial electron transfer processes in the DSC


For the sake of simplicity, only the case where the hole-transport medium is con­
stituted by the I 3 / Γ redox couple is considered here. Other hole-transport media
include organic conductors such as spiro-OMeTAD (2,2/,7,7/-tetrakis(N,N-di-p-
methoxyphenylamine)9,9/-spirobifluorene) [10.6-10.10] and inorganic p-type con­
ductors such as CuSCN [10.11, 10.12] and Cul [10.13]. Under illumination at open
circuit, the net rate of electron injection into the mesoporous oxide must be balanced
by the net rate of electron transfer to I3 ions in solution and to oxidized sensitizer
molecules (it is assumed for simplicity that the second of these processes can be
neglected due to an efficient dye regeneration by Γ). In principle, electron transfer
to I3 can occur either at the interface between the nanocrystalline oxide and the
electrolyte or at electrolyte-exposed areas of the anode contact (usually a fluorine-
doped tin oxide layer on glass). In practice, the second route can be suppressed
by using a compact blocking layer of oxide deposited on the anode through spray
pyrolysis [10.14, 10.15]. Blocking layers are mandatory for DSCs that utilize one-
electron redox systems such as cobalt complexes [10.16-10.18] or for cells using
solid organic hole-conducting media [10.6, 10.9]. In the absence of electron transfer
to I3 via the anode substrate or to oxidized dye molecules from the nanocrystalline
Characterization and modeling of dye-sensitized solar cells 325

Ti0 2 , the average steady-state electron density under illumination at open circuit is
determined by the condition

(10.1)

Here, L>inj and vhr are global volume rates for electron injection and for 'back
reaction' of electrons with I3, respectively. In the simplest case of uniform dye loading
and uniform illumination (i.e., weakly absorbed light), the rates of electron injection
and back reaction are constant throughout the film. The global rate of electron injec­
tion (cm-3 s_1) depends on η^ the fraction of the incident photon flux (70) - corrected
for reflection losses - that is absorbed by the sensitized mesoporous film (thickness d),
and on the injection efficiency, η^.

(10.2)

The rate of the back reaction

(10.3)

depends on the reactant concentrations and the rate constants for electron transfer.
Since reaction (10.3) is not an elementary step, the order of the reaction with respect
to the reactants is not easily defined a priori. In formal terms, we can write

(10.4)

where khr is the rate constant for the back reaction of electrons, and the v terms repre­
sent the reaction orders with respect to electrons and tri-iodide ions.
In principle, reaction (10.3) can occur by one of two routes. The first elemen­
tary step is electron capture by I3 (or I2 in equilibrium with I3) to form the iodine
radical ion
(10.5a)
There are two possible second steps involving I 2
(10.5b)
or
(10.5c)
If the first step (10.5a) is rate-determining, the reaction will be first-order in both
electron and I3 concentrations. If on the other hand reaction (10.5b) is rate-determining
with reaction (10.5a) close to a pre-equilibrium, the reaction orders vn and νγ will be 2
and 1, respectively. Finally, if reaction (10.5c) is rate-determining, both vn and v are 2.
Kinetic treatments based on equations (10.5a-c) introduce single rate constants for
electron transfer to I3 or Y2. The problem with this approach is that electrons can be trans­
ferred from the conduction band (free electrons) or from surface states [10.19] located in
326 Dye-Sensitized Solar Cells

the bandgap of the oxide. The rate constants for electron transfer from these two types of
states will not be the same. In order to progress towards a testable model, several simpli­
fying assumptions are made at this point. Electron transfer via surface states is neglected,
i.e., it is assumed that only conduction band electrons are involved in the back reaction
(the case where electron transfer occurs via surface states is complex). Furthermore, the
reaction orders with respect to electrons and tri-iodide will be taken as 1. Under such con­
ditions, the steady state condition, Eq. (10.1), implies that the steady-state free electron
density at open circuit under illumination is a linear function of light intensity, given by:

(10.6)

Here, r 0 = l/^brUa] is m e conduction band electron lifetime determined by the


back reaction with I3.
Direct measurements of the absolute conduction band electron density in a
DSC are difficult, although changes in conduction band density can be monitored by
measuring the conductivity of the mesoporous oxide as a function of the experimen­
tal conditions [10.20-10.22]. The most important indirect measure of the conduction
band electron density is the open circuit photovoltage, Voc. In the dark, electrons in
the oxide are in equilibrium with the Y^ßr redox system, which is characterized by its
redox Fermi energy, ^redox· The (very low) equilibrium electron density in the con­
duction band of the oxide is defined by the energy difference between the redox Fermi
energy and the conduction band energy, Ec, which is on the order of 1 eV [10.23]. If
the electrons behave ideally and are non-degenerate, their equilibrium density, nc>eq, is
given by the Boltzmann limit of the Fermi Dirac distribution.

(10.7)

where Nc is the density of conduction band states. When the DSC is illuminated, elec­
trons in the mesoporous oxide are no longer in thermodynamic equilibrium with the
redox system. Instead, an injection of electrons leads to a photostationary state where
the steady-state electron density is defined by Eq. (10.1). Under non-degenerate con­
ditions, nc, the density of free electrons in the conduction band of the oxide defines the
electron quasi-Eermi level, nEv.

(10.8)

The open circuit voltage developed by the DSC corresponds to the change in
Fermi level in the oxide (and hence in the contact) brought about by illumination as
shown in Figure 10.1. It is given by

(10.9)
Characterization and modeling of dye-sensitized solar cells 327

Fig. 10.1 The upward shift (qUphoio) of the Fermi level in the DSC under illumination cor­
responds to an increase in the density of conduction band electrons by many orders of
magnitude.

The leV energy difference between the conduction band and the I3 IV Fermi
level corresponds to a very low equilibrium electron density of ~10 3 cm - 3 in the con­
duction band. When the DSC is illuminated so as to generate a typical photovoltage
of 0.75 V, nc increases to ~10 1 6 cm - 3 .
If the assumptions made above regarding ideality and the reaction order with
respect to electron density are correct, it follows from Eq. (10.6) that the photovoltage
should be given by

(10.10)

Eq. (10.10) predicts that the photo voltage should increase by 59 mV for every
decade of intensity at 298 K. In practice, DSCs are always non-ideal to some extent,
and the photovoltage generally varies by more than 59 mV per decade of intensity
(values as high as 110 mV/decade are not uncommon). The origin of this non-ideality
has yet to be understood, but it has important consequences for the interpretation of
many of the techniques discussed in this chapter. An empirical non-ideality factor, m
(> 1), is often employed to account for non-ideality, so that the intensity dependence
of the photovoltage is given by

(10.11)

In the theoretical treatments in the following sections, it is assumed that the


DSC behaves ideally (m = 1). However, it is important to recognize that non-ideality
328 Dye-Sensitized Solar Cells

influences the interpretation of the results obtained by many of the techniques


described below since Eq. (10.7) is assumed to be valid [10.24].

10.2.2 Electron trapping in the DSC


It has become clear from studies of DSCs that the density of electrons present in
the devices at open circuit under illumination is three to four orders of magnitude
higher than expected on the basis of the discussion in the preceding section. Electrons
appear to be trapped at energy levels located in the bandgap of the oxide. The origin
of the electron traps remains obscure at present: they could correspond to trapping
of electrons at defects in the bulk or surface regions of the mesoporous oxide or to
Coulombic trapping due to interaction of electrons with the cations of the electrolyte.
It has been observed that the energy distribution, g(Et), of electron trap states can
often be described by an expression of the form [10.19, 10.25-10.31]

(10.12)

Here, Nt is the total trap density (typically 1019-1020 cm-3) and T0 is a character­
istic temperature that is generally found to be considerably higher than ambient tem­
perature (T0 = 600-1500 K). This means that the trapped electron density increases
more slowly than the conduction band electron density as the quasi Fermi level moves
up. It is generally assumed that electrons are exchanged between the trap levels and
the conduction band in such a way that the trap occupancy is determined by the Fermi
Dirac function. If the system is illuminated at open circuit or if the forward bias volt­
age is increased in the dark, the upward movement of the electron quasi-Fermi level
increases the density of trapped electrons by an amount Ant given by:

(10.13)

where fJjfht and f^k are the Fermi Dirac functions


J FD ./ FD

(10.14)

As an approximation (the 'zero Kelvin approximation'), the two Fermi Dirac


functions can be replaced by step functions at EF redox and nEF, respectively, to give
the simpler expression:

(10.15)

The change in trap occupancy is illustrated in Figure 10.2, where Ant corre­
sponds to the hatched area.
Characterization and modeling of dye-sensitized solar cells 329

Fig. 10.2 The density of states function, g(Et), for electron traps showing occupancy deter­
mined by the quasi-Fermi level, nEF, which can be modified by illumination or by applied volt­
age. Ant represents the change in trapped electron density.

It follows from this discussion that the density of trapped electrons is generally
many orders of magnitude higher than the density of electrons in the conduction
band.
An important consequence of the exchange of electrons between trap states
and the conduction band is that perturbation of a DSC (from an initial steady state
condition) changes the density of trapped electrons as well as that of electrons in the
conduction band. If, for example, we consider a small step in illumination intensity
at open circuit, the time it takes to reach a new steady state is determined by the time
required to re-establish the condition df/dt = 0, in which the rate of electron trap­
ping is balanced by the rate of thermal release of electrons from traps [10.32, 10.33]
(here, / represents the trap occupancy). This situation is schematically illustrated in
Figure 10.3 for a light step, where trapping of additional electrons moves the quasi-
Fermi level up (df/dt > 0 ) until a new steady state is reached (df/dt = 0).
As a consequence of the 'buffering' effect of traps, the measured relaxation
times associated with electron transport and with electron transfer to I3 depend on the
light intensity [10.27, 10.34, 10.35]. Bisquert and Vikhrenko [10.33] have presented
a detailed treatment of this problem in terms of the quasi-static approximation, so we
here simply note the main results of their analysis for the open circuit case. It can be
shown that the time constant, rn, for the relaxation of the conduction band electron
density is given by:

(10.16)
330 Dye-Sensitized Solar Cells

Fig. 10.3 The response of a DSC to a small illumination step. The initial steady state is per­
turbed by the increase in illumination intensity, and the rate of trapping exceeds the rate of
detrapping so that the quasi-Fermi level moves up until a new steady state is established (equal
trapping/detrapping rates). Note also that the increased density of electrons in the conduction
band increases the rate of electron transfer to I"

where r 0 = l/^brUa] (cf. Eq. (10.6)). The dnt/dnc term in Eq. (10.16) reflects the way
that the densities of trapped and free electrons vary with changes in the quasi Fermi
level (cf. Eqs. (10.12) and (10.8), respectively).

(10.17)

Eqs. (10.16) and (10.17) explain the observation of the effective electron life­
time (i.e., the time constant for relaxation of the conduction band electron density)
decreasing as the bias light intensity is raised, since nc varies more rapidly with nEF
than g(nE¥) if the characteristic temperature of the trap distribution (Γ0) is greater than
the ambient temperature (7). Figure 10.4 illustrates how the ratio TJTQ is predicted to
vary with Ec - nER or with qUphoto.
The slope of the linear part of the semi-logarithmic plot shown in Figure 10.4
is given by:

(10.18)
Characterization and modeling of dye-sensitized solar cells 331

Fig. 10.4 The effective electron lifetime, rn, decreases with increasing light intensity as the
quasi-Fermi level moves closer to the conduction band energy. Note that rn tends towards the
limiting value r0 as nEF approaches Ec. Ec- EFjedox = 1 eV.

This means that the measurement of the effective lifetime as a function of pho­
tovoltage should give values of the characteristic temperature, T0.
The time constant, rn, associated with a small perturbation of the DSC can be
obtained from the transient (or periodic) photovoltage response, since a linear expan­
sion of Eq. (10.8) for small modifications shows that the change in quasi-Fermi level
8nEF (and hence photovoltage £Uphoto) is linearly related to the change in the conduc­
tion band electron density 8nc.

(10.19)

In principle, it should be possible to determine r 0 from the relaxation time


constant of the photovoltage at very high intensities (i.e., high values of nc) when
dnt/dnc becomes much smaller than 1 (cf. Eq. (10.16)). However, this limit has not
been experimentally observed.

10.2.3 Electron transport in the DSC

The 'driving force' for the transport of electrons and ions in the DSC is the gradient of
free energy or electrochemical potential, 3μ ί /3χ, which is equivalent to the gradient
332 Dye-Sensitized Solar Cells

of the Fermi energy, dEF/dx [10.36]. The electrochemical potential^of a species /


with charge z\q is defined as

(10.20)

where φ is the inner potential, nx is the concentration of the species and n\ is the con­
centration of some reference state for which the chemical potential takes its standard
value μ°. The flux of charged species depends on the gradient of free energy

(10.21)

where u^is the mobility, and the gradient of electrochemical potential is given by:

(10.22)

It follows that the flux of species is expressed as:

(10.23)

Using the Einstein relation D{ ■ which relates the diffusion coefficient D{ to the
mobility, we obtain:

(10.24)

The first term is equivalent to Fick'sfirst law of diffusion, and the second term repre­
sents the migration or drift flux.
Electrons injected into the Ti0 2 from the photo-excited dye must move
through the network of interconnected oxide nanoparticles to reach the anode. The
excess electronic charge in the oxide arising from injection of electrons is balanced
by a net positive ionic charge in the electrolyte arising from the regeneration of
the dye by electron transfer from iodide ions. The density of ionic charges in the
electrolyte is high (~1020 cm -3 ), so that excess charges in the mesoporous oxide
are effectively shielded and electric fields are screened out over short distances.
As a consequence, it appears that electrons are collected at the anode by diffusion
[10.37-10.39] (i.e., the second term in Eq. (10.24) is negligible). The strong cou­
pling between ionic and electronic charges signifies that the diffusion process is
ambipolar [10.40-10.42]. However, the diffusion of electrons is also complicated
by trapping and detrapping processes influencing the time response of the current
to external perturbations. These processes have been modeled using Monte Carlo
methods [10.30, 10.43, 10.44], but attention is here restricted to a macroscopic
description. In the absence of macroscopic electrical fields, the transport, trapping
Characterization and modeling of dye-sensitized solar cells 333

and back reaction of electrons through the mesoporous oxide film are described by
the continuity equation [10.45]

(10.25)

Here, the first term represents the local rate of injection of electrons into the
oxide, the second term accounts for trapping and detrapping (f is the probability of
trap occupation), the third term describes the time-dependent diffusion of electrons
and the final term allows for the loss of electrons by back reaction with tri-iodide.
The angular brackets in the trapping/detrapping term indicate an average over the trap
state energies, Ετ, weighted by the trap density, g(Et).
Solutions of the continuity equation have been obtained for steady-state con­
ditions [10.37] as well as for various kinds of external perturbation [10.35, 10.46-
10.50]. Under conditions of steady illumination, df/dt is zero, leading to the second
term vanishing and an analytical solution being obtained. Figure 10.5(a) contrasts the
steady-state profiles of conduction band electrons calculated for the limiting condi­
tions of open circuit and short circuit. Figure 10.5(b) illustrates the corresponding
quasi-Fermi level profiles (cf. Eq. (10.8)). It can be seen that the quasi-Fermi level is
flat in the open circuit case, whereas in the short circuit case, it drops steeply close to
the anode.
Solutions of the continuity equation for non steady-state conditions can be
obtained numerically [10.45]. However, in early work (for example using intensity
modulated photocurrent spectroscopy [10.48, 10.51]), the trapping term was omitted,
and the continuity equation was solved using an intensity-dependent effective diffu­
sion coefficient, Dn. Experimentally, Dn has been observed to decrease with intensity
(compare this with rn, which decreases with intensity) [10.51]. The intensity depend­
ence of the diffusion coefficient arises from the relaxation of the density of trapped
electrons that are exchanged with the mobile electrons in the conduction band. The
time constant associated with this relaxation has been discussed by Bisquert and
Vikhrenko [10.33] within the framework of the quasi-static approximation that was
used in the preceding section to describe the relaxation of conduction band electron
density under open circuit conditions. Their treatment shows that the effective dif­
fusion coefficient depends on trap occupancy and hence of the Fermi level. Dn is
related to the diffusion coefficient of free electrons, D0, in the mesoporous oxide by
the expression:

(10.26)

(cf. Eq. (10.16)). Figure 10.6 illustrates the increase in Dn as the quasi-Fermi level
moves towards the conduction band. The corresponding variation of r n has been
included in the figure in order to demonstrate that the product D n r n is predicted to be
constant.
The problem with using an effective diffusion coefficient to obtain time-
dependent solutions of the continuity equation under short circuit conditions is that
334 Dye-Sensitized Solar Cells

Fig. 10.5 (a) Steady-state profiles of conduction band electron density calculated for open
circuit and short circuit conditions for a DSC with a 10-μηι Ti0 2 layer. I0 = 1017 cm -2 s-1,
a = 500 cm-1, r 0 = 10~4 s, D0 = 0.5 cm2 s_1. Illumination from the anode side (x = 0). Note the
different y scales, (b) Quasi-Fermi level profiles corresponding to the conduction band electron
density profiles shown in Figure 10.5(a).
Characterization and modeling of dye-sensitized solar cells 335

Fig. 10.6 The predicted increase of the effective electron diffusion coefficient with increasing
light intensity as the quasi Fermi level moves towards the conduction band. As nEF approaches
Ec, Dn tends towards D0 (0.5 cm2 s_1). The corresponding decrease in the apparent electron life­
time is also shown (r 0 = 10~4 s). Note that the electron diffusion length Ln = (Dnrn)m remains
constant. Ec - E^redox = 1 eV.

the quasi-Fermi level - and hence the dnc/dnt term in Eq. (10.19) - vary with distance.
Under conditions where the diffusion coefficient varies with distance, diffusion can
no longer be described by Fick's second law (the problem of diffusion in an inhomo-
geneous medium has been discussed by van Milligen et al. [10.52], who demonstrated
that an additional term will appear in the diffusion equation if the diffusion coefficient
varies with position). However, comparisons of solutions of the continuity equation
using the effective diffusion coefficient approximation with the results of the numeri­
cal solution of the time-dependent problem including the full trapping term reveal that
the approximation gives a fit that is adequate, at least for light pulse or light step meas­
urements. For this reason, the concept of an effective diffusion coefficient remains
useful for analyzing experimental data.
The preceding sections highlight the need to consider trapping and detrapping
of electrons when interpreting the results of non-steady-state experiments on DSCs.
Emphasis has been placed on the quasi-static treatment of Bisquert and Vikhrenko
since it has the merit of being testable and suitable for incorporation into detailed
models of the response of DSCs to external perturbations. However, several aspects
of the behavior of DSCs suggest that the approach may require refinement or modi­
fication in the light of new experimental results. In particular, the non-ideality factor
seen in the intensity dependence of the photovoltage appears to affect the relationship
336 Dye-Sensitized Solar Cells

between the quasi-Fermi level and the conduction band electron density, causing
Eq. (10.8) to no longer be valid [10.24].

10.3 THE TOOLBOX

10.3.1 Determination of injection efficiency and electron diffusion


length under steady-state conditions
Under steady-state conditions, the rate of trapping and detrapping should be equal in
order for the presence of trap states not to have any influence on for instance the electron
diffusion length or the charge collection efficiency. Thus, steady-state measurements
avoid some of the complications involved in finding the electron diffusion coefficient,
Ln, by the perturbation frequency domain or transient methods. A standard method for
studying various parameters under steady-state conditions consists in measuring the
spectral incident-photon-to-collected-electron efficiency (IPCE), defined as:

(10.27)

where j s c is the short circuit current density of the cell under incident monochromatic
light with wavelength A and photon flux 70, and q is the elementary charge. This can also
be referred to as the external quantum efficiency (EQE) of the device. The IPCE can be
expressed as the product of the efficiencies of three separate physical processes:
(10.28)

where τ/^λ) is the light-harvesting efficiency of the sensitized oxide layer, T7inj(A) is
the efficiency of the electron injection from the sensitizer into the oxide and T7col(A) is
the electron collection efficiency.
The ratio IPCE(A)/i7ih(A) can be further defined as the absorbed-photon-to-
collected-electron efficiency (APCE(A)) to mark the division of the IPCE into its opti­
cal and electrical parts. For state-of-the-art DSCs, APCE can be practically 100 %
[10.1, 10.2, 10.53] and IPCE approximately 85 % [10.54, 10.55] depending on reflec­
tion losses at the DSC substrate.
Two recent papers have highlighted the possibilities of employing IPCE meas­
urements to determine η^ and Ln [10.56, 10.57], based on an approach developed by
Södergren et al. [10.37]. As mentioned above, APCE(A) can be described by a func­
tion where the only unknown parameters are η^ and Ln. The ratio of APCE(A), meas­
ured for illumination from the back and front sides of the cell (in the case of using a
sandwich DSC with two transparent substrates), can be shown to depend only on Ln.
Using this Ln value, we can calculate η^ from the APCE spectra.

Optical measurements
Quantitative in-situ measurement of the light harvesting efficiency of complete dye solar
cells is complicated because of light scattering by the mesoporous oxide film and light
Characterization and modeling of dye-sensitized solar cells 337

absorption by the other cell components. The literature presents several descriptions of
the procedures for obtaining η^, see for example [10.53, 10.56-10.58]. As an example,
we refer to the work by Barnes et al. [10.57] on relatively transparent Ti0 2 films.
To take into account reflective and absorption losses in the DSC that are not
attributable to the dye/oxide system, measurements of the conducting glass, platinized
counter electrode layer, and electrolyte, are carried out. We follow the notation of
Barnes et al., and denote reflectance from the glass by R, (1 - TPt) for the absorption
due to the Pt layer and (1-Ti) for the absorption in the electrolyte between the Ti0 2
and the Pt layer. The fraction of light transmitted by complete cells with dye Crdye/Ti0 )
and without dye (ΓΉ0 ), with a mesoporous film thickness d gives the absorption coef­
ficient of the dye-coated mesoporous film according to:

(10.29)

The assumptions underlying Eq. (10.29) are that there is an exponential varia­
tion of light intensity with position in the mesoporous film and that the dye molecules
do not scatter a significant fraction of light, i.e., TàydTi0/TTi0i - [r dye/Ti0 /(l - Rdy&mÖ2)l
The fact that iodine in the electrolyte, filling the oxide pores, absorbs light, needs to be
taken into account, which is done by determining its absorption coefficient, α^λ), by
measuring transmission with and without iodine in the electrolyte and by estimating the
porosity of the films. Barnes et al. [10.57] utilized relatively non-scattering Ti0 2 films
and scattering effects were negligible for Λ > 480 nm. To describe scattering effects, the
optical measurements should be performed with an integrating sphere, and more sophis­
ticated models, such as Kubelka-Munk [10.59], which include the scattering properties,
would be more appropriate.
To obtain η^, we need to know the light absorption profile throughout the dyed
mesoporous oxide film. The generation rate of excited dye states as a function of the
position x in the film is approximated by
(10.30)

for illumination from the photoelectrode (PE), where x = 0 is the position of the FTO
substrate-Ti02 interface. Illumination from the counter electrode-electrolyte side
(CE) gives
(10.31)

Integration of Eqs. (10.30) and (10.31) across the thickness of the oxide film provides
the light-harvesting efficiency for PE and CE illumination [10.57]

(10.32)

(10.33)
338 Dye-Sensitized Solar Cells

Fig. 10.7 Light-harvesting efficiencies for a compressed Ti02 (P25) DSC adapted from ref.
[10.56]. Thefilmswere heated after compressions at 450 °C for 1 h. The dye was N719 and
the electrolyte consisted of 0.6 M hexylmethylimidazodiumiodide, 0.05 M I2 and 0.5 M tert-
butylpyridine in 3-methoxypropionitrile. (a) PE illumination and CE illumination, (b) The
dependence of the light harvesting on the Ti02 thickness.

Figures 10.7(a) and 10.7(b) are taken from ref. [10.56], and show (a) η^ for
a DSC illuminated from the different sides and (b) as a function of the film thick­
ness. The solar cell consisted of a mesoporous Ti0 2 film prepared by compression
of Degussa P25 nanoparticles, N719 dye and the iodide/tri-iodide redox couple in a
3-methoxyproprionitrile solvent. It should be noted that a film based on P25 nanopar­
ticles scatters light significantly and scattering effects were carefully considered in
this case -the reader is referred to the original paper [10.56] for details.
Optical losses limit the maximum η^ to about 86 % and 79 % at the PE and CE
illumination, respectively. In Figure 10.7(b), we note that at the dye absorption peak
(535 nm) η^ reached its maximum at a film thickness of 8-9 μιη, whereas for weak
absorbing light with Λ > 700 nm, r/ih continued to increase even for d > 30 μιη.

IPCE measurements
Incident photon conversion efficiencies are typically obtained using a xenon or hal­
ogen lamp coupled to a monochromator. The photon flux of incident light on the
samples is measured with a calibrated photodiode, and measurements are typically
carried out at 10- or 20-nm wavelength intervals between 400 nm and the absorption
threshold of the dye. Since the DSC is a device with relatively slow relaxation times,
it is important to ensure that the measurement duration for a given wavelength is suf­
ficient for the current to be stabilized (normally 5-10 seconds). If IPCE is observed to
depend on the light intensity, the measurements should be performed with additional
bias light to ascertain that IPCE is determined at relevant light intensity conditions.
The reasons for light-intensity-dependent IPCE may be that T7col increases with light
intensity due to faster electron transport, or that there are mass transport limitations in
the electrolyte, decreasing IPCE with light intensity.
Characterization and modeling of dye-sensitized solar cells 339

The compressed Ti0 2 (P25) films studied by Halme et al. showed relatively low
IPCE values [10.56]. The IPCE spectra for both PE and CE illumination are shown
in Figure 10.8.
The T7IPCE peak near 540 nm corresponds to the absorption maximum of the dyed
Ti0 2 films. A distinctive feature is that the IPCE is relatively low in this study. As men­
tioned above, IPCE values were above 80 % for the best DSC. The maximum IPCE was
ca. 40 % for PE illumination and approximately 25 % for CE illumination, compared
to the maximum ^h of ca. 83 % in this study. This proves that these cells suffer from
substantially low T7col values, and hence low APCE. The effect of the reversed direction
of illumination is also shown in Figure 10.8. For strongly absorbed wavelengths, IPCE
decreased rapidly for CE illumination. As a consequence, the IPCE, CE peak shifted
towards longer wavelengths. For high a, most of the light is absorbed in the outer layers
in the Ti0 2 film. The electrons thus have a longer distance to move, as compared to PE
illumination with high a, before they are collected at the back contact. Consequently, they
are more prone to recombination losses. Quantitatively, we expect an effective electron
diffusion length in the same range as the film thickness.

Estimations ofLn and η^


Comparison of spectral IPCE measurements taken for opposite illumination direc­
tions is not only a good diagnostic test for detecting low T7col but also provides a basis
for quantitative estimations of Ln by the diffusion model. The derivation can be found
in refs. [10.56, 10.57].
The standard diffusion model of electron generation, transport and recombina­
tion in nanostructured photoelectrodes is based on the continuity equation for electron
concentration [10.37]:

(10.34)

Fig. 10.8 IPCE spectra of compressed Ti02 (P25) DSC reported in ref. [10.56]. The films
were heated after compressions at 450 °C for 1 h. The dye was N719 and the electrolyte con­
sisted of 0.6 M hexylmethylimidazodiumiodide, 0.05 M I2 and 0.5 M tert-butylpyridine in
3-methoxypropionitrile. (a) PE illumination, (b) CE illumination.
340 Dye-Sensitized Solar Cells

where D and r are, respectively, the electron diffusion coefficient and lifetime, n the
local electron density, n0 the equilibrium electron density in the dark, and G the local
electron generation rate according to Eqs. (10.30) and (10.31). The position coordi­
nate x = 0 is located at the film edge facing the in-coming light.
When the light is incident on the cell from the PE side, the boundary conditions
for solving Eq. (10.34) at the short circuit condition are:

(10.35)

(10.36)

For CE illumination, x = 0 at the counter-electrode-facing edge of the film, and the


boundary conditions are:
(10.37)

(10.38)

The model assumes that the electron transport occurs via diffusion, that the
recombination reactions are first-order in the electron concentration, that D and r are
independent of x and n(x), and that under short circuit conditions, extraction of elec­
trons at the substrate contact is fast enough to maintain the excess electron concentra­
tion at the contact close to the dark equilibrium value.
By solving Eqs. (10.34-10.38), and substituting Ln = (£>r)1/2, the ratio between
the IPCE, or APCE, spectra, which depends only on α(Λ), d and Ln, since T7inj(A) can
be assumed independent of the direction of illumination can be written as:

(10.39)

Two special cases can be considered with respect to uniform or highly non-uni­
form electron generation. For uniform generation, obtained in the limit of weak light
absorption, T7col becomes

(10.40)

independently of the illumination direction. In the opposite case lia « d, corre­


sponding to high T7lh, T7col approaches 100 % for the PE illumination, whereas for the
CE illumination, it becomes

(10.41)
Characterization and modeling of dye-sensitized solar cells 341

The main predictions from the above equations can be summarized as [10.56]:

1. For constant Ln and a, ηοο{ decreases with d for the both directions of
illumination.

2. For Ln much larger than d, η00γ approaches 100 %, irrespective of a and the
direction of illumination.

3. For uniform electron generation, obtained in the limit of weak light absorp­
tion, T7col becomes equal at the both illumination directions, irrespective of d
and Ln.

4. For constant d and Ln, T7col increases with a at the PE illumination, whereas at
the CE illumination, the trend is opposite.

Using the experimental data for APCE,CE/APCE,PE, d and a, Ln can be esti­


mated directly from Eq. (10.39) for each Λ and d. With the estimated Ln, i7coican be
calculated from Eqs. (10.40-10.41), and T7inj subsequently becomes:

(10.42)

It should be noted that a fit to the IPCE ratio, Eq. (10.39), is independent of
calibration errors in the IPCE measurement (the ratio can be normalized to unity at
long wavelengths if the losses due to Pt and iodine are uncertain) [10.57].

Recent results using IPCE measurements to determine Ln and η^


Both the work of Halme et al. [10.56] and Barnes et al. [10.57] successfully demon­
strate the use of optical and IPCE measurements to determine η^ and Ln for various
types of DSC devices. The parameters correlated with device performance; relatively
low T7inj values and Ln ~ d resulted in moderate IPCE below 50 %, and an increase in
the diffusion length and injection efficiency led to higher short circuit current. The
diffusion length was found to depend on light intensity and film thickness, i.e., Ln
depended on the electron concentration and was thus not constant throughout the mes-
oporous Ti0 2 film. A possible reason for this is surface-trap-mediated recombination.
Another way to take this effect into account would be to solve the diffusion equation
in which Ln explicitly depends on the electron concentration. The dependence of Ln
on electron concentration needs further research, both in terms of developing new, and
predictive, models, and relevant experimental characterization methods.
Another intriguing finding in refs. [10.56, 10.57, 10.60] was that Ln determined
by IPCE measurements was two to three times shorter than what has been found
by transient methods. This significant discrepancy is at present poorly understood.
Barnes et al. [10.57] discussed it as being related to the measurement of Dn with tran­
sient methods at short circuit, whereas Jennings [10.60] suggested that the injection
efficiency may be a function of position in the mesoporous film due to a non-uniform
distribution of non-injecting dye aggregates. The fact that Ln varies to such an extent
depending on which method is used to determine it needs to be resolved, since the
342 Dye-Sensitized Solar Cells

value of the diffusion length has very important implications to the performance opti­
mization of DSCs.

10.3.2 Electrochemical and spectrolectrochemical techniques to study


the energetics of the oxide/dye/electrolyte interface
The position of the energy levels at the oxide/dye/electrolyte interface are fundamen­
tally important to the function of the DSC. This is conventionally drawn as a sche­
matic energy level diagram as depicted in Figure 10.9, taken from ref. [10.61]. The
desired pathway for a photoexcited electron is indicated, as are the potentials for a
conventional DSC based on the RuL2(SCN)2 dye adsorbed on Ti0 2 and with I/I3 as
the redox couple in the electrolyte.
Standard measurements of energy levels for semiconductors, redox couples and
dyes in solutions or adsorbed to surfaces can be found in many textbooks (see for
example refs. [10.62] and [10.63]) and will not be described in this section. Rather,
we focus on specific electrochemical and spectroelectrochemical experiments within
the DSC tool-box approach, i.e., to obtain data of the energy levels at the oxide/dye/
electrolyte interface from in situ measurements.
Care needs to be taken when using an energy level diagram such as the one
depicted in Figure 10.9 for interpretation and analyses of the actual energetic situation

Fig. 10.9 A schematic representation of the principle of the DSC to indicate the electron energy
levels in the various phases. The potential values are indicative of a DSC consisting of the RuL2
(SCN)2 dye adsorbed on Ti02 and I7I~ as the redox couple in the electrolyte. The cell voltage
observed under illumination corresponds to the difference, AV, between the quasi-Fermi level
of Ti02 and the electrochemical potential of the electrolyte. S stands for sensitizer; S*, elec­
tronically excited sensitizer; S+, oxidized sensitizer. Adapted from Figure 9 of ref. [10.61].
Characterization and modeling of dye-sensitized solar cells 343

in a DSC. The indicated values may be obtained from measurements of the individual
components, excluding effects due to adsorption of the dye, and electrolyte species,
to the oxide surface. The ordinate in Figure 10.9 may present internal energy and not
free energy. The charge carriers present a significant configurational entropy arising
from the number of accessible energy states, which can differ significantly in the vari­
ous phases. Sometimes the energy levels of the dye are reported as HOMO/LUMO
levels, obtained for example from electronic structure calculations involving several
approximations.

A matter of scale
Figure 10.9 also represents an 'interface' between physics and chemistry. In solid-
state physics and electrochemistry, one normally uses the energy scale with vacuum
as reference for the former and the potential scale with the standard hydrogen elec­
trode (SHE) or normal hydrogen electrode (NHE) as reference for the latter. The
electrochemical potential of electrons in a semiconductor is normally referred to as
the Fermi level, £F, whereas in an electrolyte solution it is generally denoted the redox
potential. At equilibrium, the electrochemical potentials (or the Fermi levels) of the
semiconductor and electrolyte are equal. For most purposes in electrochemistry, it is
sufficient to reference the redox potentials to the NHE (or any other, more practical,
reference system), but it is sometimes of interest to have an estimate of the absolute
potential (i.e., versus the potential of a free electron in vacuum). For example, it may
be interesting to estimate relative potentials of semiconductors, redox electrolytes
and solid hole-conductors in solid-state DSC based on their work functions. Taking a
redox system dissolved in an electrolyte as an example, the absolute energy of a sys­
tem is shifted against the conventional scale by the free energy, Erei, according to:

(10.43)

in which £p redox is the standard redox energy level (or potential, depending on which
unit is the most appropriate to use). For the NHE, £refis estimated to be -4.6 ±0.1 eV
[10.63]. With this value, the standard potentials of other redox couples can be
expressed on the absolute scale.

Energy levels in semiconductors


In non-degenerate semiconductors, the equilibrium Fermi level is given by:

(10.44)

where Ec is the energy at the conduction band edge, kBT is the thermal energy, nc is
the density of conduction band electrons, and Nc is the effective density of conduction
band states. With respect to vacuum, Ec is given by the electron affinity EA, as shown
in Figure 10.10. The ionization energy, /, in the same figure, determines the position
of the valence band EY, whereas the distance between the vacuum level and the equi­
librium Fermi level is the work function, φ. Thus, the positions of the energy bands
can be predicted from electron affinity values.
344 Dye-Sensitized Solar Cells

Fig. 10.10 Positions of energy bands for a semiconductor with respect to the vacuum level.

However, these values are very sensitive to the environment, and measurements
of absolute and relative energies in vacuum must be carefully interpreted and analyzed
in terms of their relevance to DSCs. In the field of semiconductor electrochemistry, the
standard approach of determining the so-called flatband potential of a semiconductor,
Vfb, which estimates the work function of the semiconductor in contact with the spe­
cific electrolyte, is Mott-Schottky analysis of capacitance data [10.63]. This approach
is based on the potential-dependent capacitance of a depletion layer at the semicon­
ductor surface. For a DSC, one would not expect such behavior to be observed for
the -20 nm anatase nanocrystals that are thought to be fully depleted - see ref [10.64]
and the references therein. Instead, cyclic voltammetry and spectroelectrochemical
procedures have been used to estimate Ec. These techniques also provide information
on the density of states (DOS) of the semiconductor.
For a recent review on the measurements of Ec, and the density of states in
mesoporous Ti0 2 films, we refer to ref. [10.64]. Fitzmaurice has reviewed the first
spectroelectrochemical measurements of Ec for transparent mesoporous Ti0 2 elec­
trodes [10.65]. Using an accumulation-layer model to describe the potential distribu­
tion within the Ti0 2 nanoparticle at negative potentials [10.66], and assuming that
Ec remains fixed as the Fermi level is raised into accumulation conditions, Ec values
in organic and aqueous electrolytes were estimated. At present, there is an extensive
compilation of data showing that Ec is not that well-defined. Many electrochemical
and spectroelectrochemical studies indicate that mesoporous Ti0 2 films possess a tail­
ing of the DOS (trap states) rather than an abrupt on-set from an ideal Ec. In section
10.3.7, we present several methods for determining the density of trapped electrons
in DSCs. Nevertheless, the Ec values estimated in the early work by Fitzmaurice and
coworkers are still used today in order to, at least qualitatively, discuss for example
the energy level matching between the conduction band edge of the oxide and the
excited state of the dye.
The cyclic voltammogram of a mesoporous Ti0 2 film in an aqueous electro­
lyte is shown in Figure 10.11(a) [10.67]. The large reversible peak is indicative of
the filling and emptying of the Ti0 2 DOS, whereas the smaller cathodic pre-peak
is assigned to the filling of deep trap states. Thus, cyclic voltammetry with the use
of a reference electrode provides indications of the position of the energy levels in
Characterization and modeling of dye-sensitized solar cells 345

Fig. 10.11 (a) Cyclic voltammogram of a nanostructured Ti0 2 electrode in aqueous


LÎC104 (0.2 M), pH 6.2. Scan rate 5 mV s"1. Taken from Figure 3 in ref. [10.67]. (b)
Spectroelectrochemistry of lithium ion insertion in a mesoporous Ti0 2 electrode. The UV-Vis
spectrum changes during electrochromic switching from being essentially colorless at a reverse
bias of-0.64 V vs. Ag/AgCl (sat. KC1 in water) to showing an intense blue color following Li+
insertion at a forward bias of-1.64 V. Redrawn with data from Figure 12 in ref. [10.61].

the semiconductor, although the determination of for example Ec, estimated from the
on-set of the cathodic current, is rather inaccurate. The electrons in the T i 0 2 inferred
from electrochemical measurements also demonstrate spectroscopic signatures. As
the band gap energy of anatase T i 0 2 is 3.2 eV, its ground state UV-Vis absorption
spectrum shows a threshold at 385 nm. At a reverse bias applied to the mesoporous
T i 0 2 in a spectroelectrochemical cell, there should thus be essentially no coloration
in the visible part of the spectrum. The example in Figure 10.11(b) is taken from an
346 Dye-Sensitized Solar Cells

early study of lithium ion insertion of mesoporous Ti0 2 for electrochromic applica­
tions [10.68]. The small attenuation of light at this potential (-0.64 V vs. Ag/AgCl)
was due to light scattering. Changing the potential to -1.64 V in a stepwise fashion
led to the intercalation of Li+ and the appearance of an intensely dark blue coloration
(Ti3+ states), as shown in Figure 10.11(b).
In the original work of Fitzmaurice and coworkers [10.65], the Ec of Ti0 2 was
estimated from the absorption changes as a function of the applied potential in the
same way as shown in Figure 10.11(b). The electrons in the mesoporous Ti0 2 film,
in the conduction band or in trap states, have been studied to a great extent since
the early work in the beginning of the 1990's. For further reading on the various
experimental and theoretical methods applied for these investigations as well as on
the electronic and optical properties of electrons in Ti0 2 , we recommend ref. [10.64]
and the references therein. Spectroelectrochemical studies as described above have
also been applied to other mesoporous semiconductor electrodes such as ZnO [10.69,
10.70] and NiO [10.71]. It can be noted that, based on the above-mentioned studies,
mesoporous Ti0 2 films have been developed for electrochromic display applications
using surface-attached viologens as chromophores [10.72] and for secondary Li+ bat­
teries [10.73, 10.74].
The position of the conduction band edge depends on the surface charge (dipole
potential). The pH dependence of Ec for mesoporous Ti0 2 films in aqueous solutions
follows a Nernstian behavior with a shift of 59 mV/pH unit, due to protonation/depro-
tonation of surface titanol groups on Ti0 2 - see ref. [10.66, 10.64] and the references
therein. In non-aqueous solutions, Ec can be widely tuned by the presence of cations,
and this affect is the greatest with cations possessing a large charge-to-radius ratio.
For example, Ec has been reported to be -1.0 V vs. SCE in 0.1 M LiC104 acetonitrile
electrolyte and —2.0 V when Li+ was replaced by TBA+. The large variation of Ec
with various 'potential-determining' ions can be explained by cation-coupled reduc­
tion potentials for Ti0 2 acceptor states, due to surface adsorption and/or insertion into
the anatase lattice [10.64]. This cation-dependent shift in Ec is used to promote pho-
toinduced electron injection from the surface-bound sensitizer. For this to occur, Ec
must be at a lower energy than the excited state of the sensitizer, S+/S*. In contrast, one
would like Ec to be at as high an energy as possible to achieve a high photovoltage, see
Figure 10.9. One thus needs to compromise with regard to the position of Ec in order
to attain an efficient electron injection while simultaneously maintaining a high pho­
tovoltage. Additives in the electrolyte are normally used to fine-tune the energy level
matching of Ec and S+/S*. The effect of the additive, most often 4-tert-butylpyridine
(4TBP), can be studied by measuring shifts of Ec depending on surface charge, and
by determining the electron lifetime. The shift of Ec is measured by charge extraction
methods as described in detail in a later section and the determination of electron life­
times is presented in the section on "Intensity-modulated photovoltage/photocurrent
spectroscopy (IMVS/IMPS)."
The addition of 4TBP to the redox electrolyte gives a significant improvement
of DSC efficiencies, mainly because of an enhancement of the open-circuit photovolt-
age, Voc. As detailed in ref. [10.75], Voc increased by 0.26 V upon addition of 0.5 M
4TBP in the electrolyte. This effect may be due to (i) a higher Ec since 4TBP is a weak
base rendering the surface charge more negative when adsorbed on the Ti0 2 , (ii) a
Characterization and modeling of dye-sensitized solar cells 347

higher concentration of electrons in the Ti0 2 , nc, which will be reflected by a longer
electron lifetime in the Ti0 2 . The reason for this is that 4ΤΒΡ may block the reaction
of Ti0 2 electrons with electrolyte species, (iii) a change in redox potential due to
interactions with 4TBP and the redox species in the electrolyte. The charge extrac­
tion method was used to study the relation between charge and Voc without and with
addition of 4TBP: see Figure 10.12(a) [10.75]. Electron lifetimes were measured by
the small amplitude photo voltage transient method at open-circuit conditions. Figure
10.12(b) shows electron lifetimes as functions of the extracted charge for electrolytes
with and without 4TBP.
The analysis of the data in Figure 10.12 was as follows: The potential of the
Ti0 2 electrode was given by the quasi-Fermi level, nEF. The measured photovoltage
in the DSC was provided by the difference between nEv and the redox potential of
the electrolyte. No significant change in the redox potential occurred when 4TBP
was added to the electrolyte, and alternative (iii) above could thus be ignored. The
charge extraction results in Figure 10.12(a) demonstrated that the addition of 4TBP
gave rise to a shift of the electronic states in Ti0 2 toward higher energies. At the same
total electron concentration, n, in the porous Ti0 2 film, the developed voltage was
higher by about 0.16 V when 4TBP was present. Assuming that the energy levels of
the trap states shifted as much as Ec with the surface charge, and that Nc remained
constant with the addition of 4TBP, we can conclude from Eq. (10.44) that Ec was
shifted 0.16 eV towards higher energies with the addition of 4TBP. The additional
0.1 V required to explain the increase of 0.26 V in Voc with the presence of 4TBP in
the electrolyte was then attributed to longer lifetimes of the Ti0 2 electrons, and thus a
higher concentration, under open-circuit conditions.

Fig. 10.12 (a) The extracted charge as a function of the open-circuit potential in a Ti02-based
DSC. (b) Electron lifetimes as functions of the extracted charge. The dotted lines correspond
to power-law fits. The 4TBP concentration in the electrolyte is indicated. The N719 dye, and
an electrolyte consisting of 0.7 M Lil and 0.05 M I2 in 3-methoxyproprionitrile were used.
Redrawn with data from Figures 2 and 3 of ref. [10.75].
348 Dye-Sensitized Solar Cells

Energy levels of redox systems in solution


The electrochemical potential of electrons, or the Fermi level, for a 1-electron redox
couple is given by the Nernst equation and can be written as [10.76]:

(10.45)

where cox and cred are the concentrations of respectively the oxidized and reduced
species of the redox system. Besides the Fermi energy, we also need a description of
the energy states being empty or occupied by electrons. The electronic energies of a
redox system are presented in Figure 10.13 and are based on the model developed by
Genscher [10.63, 10.77, 10.78].
In this energy scale, £r°ed corresponds to the energy position of occupied elec­
tron states and E®x to the empty states. They differ from the Fermi level E^ Tedox by the
so-called reorganization energy, Λ. This reorganization energy is the energy involved
in the relaxation process of the solvation shell around the reduced species following
transfer of an electron to the vacuum level. For the reverse process, i.e., electron trans­
fer from vacuum to the oxidized species, there is an analogous relaxation process. It
is normally assumed that Λ is equal for both processes. The electron states of a redox
system are not discrete energy levels, but rather distributed over a certain energy range
due to fluctuations in the solvation shell surrounding the molecule. This is indicated
by the distribution of energy states around £r°ed and £° x , Figure 10.13(b). Dred is the
density of occupied states (in relative units) represented by the reduced component of
the redox system, and Dox is the density of empty states represented by the oxidized
component. Assuming a harmonic oscillation of the solvation shell, the distribution
curves, Dred and Dox, are described by Gaussian functions:

(10.46)

Fig. 10.13 (a) Electron energies of a redox system using vacuum as a reference level. £r°ed =
occupied states, E®x = empty states, E®ïedox = Fermi level of the redox couple, (b) The corre­
sponding distribution functions. Adapted from Figure 4 in ref. [10.76].
Characterization and modeling of dye-sensitized solar cells 349

(10.47)

Dr°ed and D®x are normalizing factors such that/ ^ D(E)dE = 1. The half-width of the
distribution curves is given by:

(10.48)

Accordingly, the widths of the distribution function depend on the reorganiza­


tion energy, which is of importance for the kinetics of electron transfer processes at
the oxide/dye/electrolyte interface. Typical values of Λ lie within the range from a
few tenths of an eV up to 2 eV. In Figure 10.13(b), the concentration of reduced and
oxidized species were equal (Dred = Dox). Changing the concentration ratio varies the
redox potential (£ρΓ6αοχ) according to the Nernst equation, which can be graphically
illustrated as shown in Figure 10.14.

Energy levels of excited molecules


In the case of dyes, several oxidation states exist. Each state can be qualitatively cor­
related with the energy of molecular orbitals, as shown in Figure 10.15.
The reduction of a sensitizer molecule, S, occurs by electron transfer from an elec­
tron donor to an unoccupied level of S, thus giving the reduction potential of the molecule
with a Fermi level E^(S~/S). The oxidation of S occurs by an electron transfer from the
lower-lying occupied state to a suitable acceptor molecule. The Fermi level of the oxida­
tion potential of S is E^S/SP), and the difference between these two redox potentials can
only roughly be taken as the difference of the lowest unoccupied and highest occupied
states in the molecule. An excited sensitizer is more easily reduced or oxidized, due to the
excitation energy AE* stored in the molecule. The possible redox reactions are

(10.49)

Fig. 10.14 Energy states and the redox potential (E®Îedox) of a redox system for varying concen­
trations. Adapted from Figure 10 in ref. [10.76].
350 Dye-Sensitized Solar Cells

Fig. 10.15 A molecular energy scheme of molecules in (a) the ground state and (b) the excited
state [10.76].

and
(10.50)

It is now possible to estimate the redox potentials of excited molecules by adding or


subtracting AE* from the redox potential of the molecule in the ground state. The stored
excitation energy AE* corresponds to the energy of the 0-0 transition between the lowest
vibrational levels in the ground and excited states, i.e., AE* = AE0.0. We thus obtain

(10.51)

(10.52)

From these equations, it may be illustrative to schematically indicate the rela­


tive positions of the Fermi levels of a redox system in its ground and excited states,
Figure 10.16.
Introducing the corresponding distribution functions of the occupied and empty
states for the most relevant reaction in DSC, i.e., Eqs. (10.49-10.50), we arrive at a
more complete energy level diagram as compared to that of Figure 10.9 and obtain
the so-called Genscher diagram for an excited-state electron injection from surface-
bound sensitizers into the DOS of the mesoporous Ti0 2 film, Figure 10.17.
In Figure 10.17, the distribution functions of the empty and occupied states for
the ground and excited states are drawn with equal areas indicating that the concen­
trations of the various species are the same. Differences in concentration give rise to
different Fermi levels, ^ Fredox (S/S+) and EFrQdox(S*/S+), and thus varying driving forces
for electron injection and for regeneration of the oxidized dyes by the electrolyte. The
actual concentrations of the various species indicated in Figure 10.17 in a DSC device
depend on several factors, such as the Fermi levels of the semiconductor and electro­
lyte, the dye loading, the extinction coefficients and the light intensity.
Since the sensitizer is adsorbed at the oxide surface in a DSC, the measure­
ments to achieve the energy levels as indicated in Figure 10.17 should be made in situ.
Characterization and modeling of dye-sensitized solar cells 351

Fig. 10.16 A schematic diagram of the relative positions of the Fermi levels of a redox system
in its ground and excited states [10.76].

Fig. 10.17 Energy states and optical excitation energy for excitation of the sensitizer ground
state, S, to the excited state, S*, followed by electron injection to the DOS of the mesoporous
Ti0 2 film (oxidation of the excited sensitizer). Adapted from Figure 13 in ref. [10.64] and
Figure 15 in ref. [10.76].
352 Dye-Sensitized Solar Cells

In other words, the oxidation potential of S and the excitation energy AE0_0 should be
determined for the sensitizer anchored to the semiconductor. Previous studies have
shown that molecules anchored to mesoporous Ti0 2 , Z r 0 2 , or A l 2 0 3 films can be
reversibly oxidized in standard electrochemical cells provided that the surface cover­
age exceeds a percolation threshold [10.79-10.81]. Such a mechanism was presented
by Bonhôte et al. for phosphonated triarylamines adsorbed on mesoporous Ti0 2 ,
which displayed reversible electrochemistry as well as an electrochromic behavior
[10.79]. In the review of Ardo and Meyer [10.64], studies of this lateral hole-hopping
process through the adsorbed molecular layer in mesoporous films are summarized.
Here, we present just one example using a polyene-diphenylaniline dye, D5, adsorbed
onto a mesoporous T i 0 2 electrode [10.82]. A typical cyclic voltammogram is shown
in Figure 10.18.
A reversible oxidation of the D5 was found around +0.45 V vs. Fc + /Fc, thus
demonstrating the occurrence of hole conductivity in the system. As this redox poten­
tial, EF redox (S/S+), lies much positive of Ec for Ti0 2 , the charge has to move within
the monolayer of the dye through a hole-hopping mechanism as demonstrated in ref
[10.79].
What is left now in order to obtain the Genscher diagram in Figure 10.17, is
the ^p,redox(S*/S+) value. For this we need to measure the excitation energy AE0-o
for the sensitizer adsorbed to the oxide surface. This energy can be estimated by the
photoluminescence (PL) on-set, or from the intersection of the absorption and PL
spectra. For the D5 dye, AE 0 0 n a s been estimated to be 2.37 eV. If difficulties arise

Fig. 10.18 A cyclic voltammogram of D5 adsorbed onto a mesoporous Ti0 2 electrode.


Conditions: electrolyte 0.1 M (TBA)PF6 in acetonitrile, scan rate 200 mV/s. The inset shows
the molecular structure of the dye. Adapted from Figure 4 in ref. [10.82].
Characterization and modeling of dye-sensitized solar cells 353

with regard to measuring the PL spectrum, another way to estimate AE0_0 is from
the absorption on-set of the dyes adsorbed on the oxide at a certain percentage (e.g.,
10 %) of the full amplitude at the absorption maximum. The reorganization energy for
the D5 dye has been estimated by quantum chemical calculations to 0.97 eV [10.82].
Such estimates for dyes should of course be treated with care, since these molecules
are strongly adsorbed at the electrode. For the classical RuL2(NCS)2 (N3) dye, the
reported values are £ρ,Γ6άοχ (S/S+) = 0.85 V (vs. SCE) in acetonitrile, Δ£0-ο = 1-75 eV
giving ^F.redox (S*/S+) = -0.90 V vs. SCE [10.83]. The reorganization energy for N3 in
the ground state has been estimated to 0.35 eV [10.84]. The importance of the reor­
ganization energy of the different redox species in a DSC device is presently not well
understood. For the ultrafast electron injection process from the photoexcited dye to
the semiconductor, which occurs on the femtosecond time scale, there may simply be
no time for reorganization of the solvation shell. The effects of surface adsorption and
other molecular interactions also need to be taken into account.

Measurements of the recombination reaction between photoexcited injected


electrons and redox species in the electrolyte
The final part of this section aims at further illustrating the utilization of electro­
chemical tool-box techniques with very recent and intriguing findings about the
recombination reaction of electrons in mesoporous Ti0 2 with redox species in the
electrolyte. O'Regan and coworkers have shown that many of the dyes used for DSC
can play a role in promoting electron recombination with "iodine" in the electrolyte
[10.85, 10.86]. It is presently unclear whether the first electron-transfer step in the
recombination process involves the reduction of iodine or tri-iodide (which is formed
when iodine and iodide are mixed). O'Regan et al. therefore employed the term
"iodine" in their discussions about the recombination process with the understanding
that it referred to iodine or possibly tri-iodide. The tool-box techniques for analyz­
ing the function of the dye in promoting the recombination reaction are similar to
the ones described above in the discussion of the effects of 4TBP in the electrolyte.
Phtalocyanine dyes [10.85] were shown to lower the open-circuit photo voltage in
DSCs. From charge extraction measurements, it was concluded that the position of
the conduction band edge, Ec, was the same for the various phtalocyanine dyes as
well as for the compared N719 dye. In the absence of a decrease in Ec, the loss of
photovoltage was caused by an increase in recombination, which was also observed
from small perturbation photo voltage transient decays. The authors proposed that
the studied phtalocyanine dyes accelerated recombination by providing a binding
site for iodine near the Ti0 2 surface [10.85]. Further evidence for the importance
of such dye-iodine binding sites is presented in ref. [10.86], which reports that a
change of only two oxygen atoms in a heteroleptic Ru-complex, having in total 113
atoms, to two sulfur atoms in equivalent positions, resulted in a 2-fold increase in the
recombination rate and a 20-30 mV loss in Voc. It is known from the literature that,
for instance, the binding constant between iodine and sulfur in ethylthioether is sub­
stantially higher than iodine and oxygen in ethylether. The iodine-dye complexation
and its role in DSC devices is now an important field of research in order to obtain
further understanding and improvement of DSCs.
354 Dye-Sensitized Solar Cells

10.3.3 Electrochemical measurements with thin layer cells


Electrochemical measurements can be used to characterize a number of processes
that are important in the operation of DSCs. These include ionic transport in the redox
electrolyte, electron transfer at the platinized cathode and electron transfer via the
anode substrate (with or without a compact blocking layer).
The conventional DSC relies on rapid electron transfer to I3 at the cathode.
This is achieved by using platinized electrodes prepared by sputtering or chemical
deposition [10.49, 10.87-10.90]. In the case of cells illuminated through the cathode,
it is important to optimize the tradeoff between catalytic activity and electrode trans­
parency by carefully choosing the platinum loading. In the DSC, the electron transfer
processes in the mesoporous layer and at the cathode are coupled by ion transfer
through the electrolyte. Coupling of mass transport [10.91] and electron transfer is a
common feature of electrochemical systems, and the geometry of the DSC resembles
that of thin-layer cells employed in electrochemical studies. A simple way of charac­
terizing coupled mass transport and electron transfer in the DSC cell configuration is
to fabricate cells with two platinized electrodes identical to the cathodes used in the
DSC [10.18]. The electrolyte gap (Lgap) can be controlled by utilizing the same ther­
moplastic spacer material as that in the DSC. The cell is then characterized by linear
sweep voltammetry at a sufficiently slow scan rate (< 10 mV s_1) in order to allow the
steady-state current to be measured.
Under steady-state conditions in the thin layer cell, the fluxes of ions to the
cathode and anode depend on the concentration gradients of I3 and Γ, respectively.
Since the concentration profiles are linear under steady state conditions (i.e., constant
gradient), the corresponding current densities are given by

(10.53)

The limiting current density, jiim, that can be obtained in the thin layer cell con­
figuration is determined by three factors: the concentration of I3 (the concentration of
Γ is usually in tenfold excess), the anode-cathode separation, Lgap, and the diffusion
coefficient of I3. Figure 10.19 illustrates the concentration gradients of I3 and Γ in the
limiting current condition, where the concentration of I3 at the cathode is driven close
to zero. It can be seen from the figure that the limiting current density is given by

(10.54)

A measurement of j l i m therefore provides a convenient method for determining the


diffusion coefficient of tri-iodide or other redox mediators.
The current voltage behavior of the anode and cathode in a thin-layer cell (and of
the cathode in the DSC) is described by the Butler Volmer equation [10.18, 10.92]:

(10.55)
Characterization and modeling of dye-sensitized solar cells 355

Fig. 10.19 Steady-state concentration profiles for Γ and I~ in a thin-layer cell under limiting
current conditions. Since the concentration of I" is higher than that of I~, the current is limited
by the diffusion of I~, which is the minority component in the electrolyte.

Here, η is the overpotential (the difference between the applied potential and
the equilibrium potential) and j 0 is the exchange current density for the tri-iodide/
iodide couple, given by

(10.56)

where a is the transfer coefficient, and k° is the standard heterogeneous rate


constant for the tri-iodide/iodide system.
The IV characteristic of thin-layer cells with identical platinized electrodes is
determined by the anodic and cathodic overpotentials as a function of the cell current
[10.18]. Figure 10.20 illustrates how the rate constant for electron transfer at the cathode
affects the current voltage curve that would be observed for an electrode in a conventional
3-electrode setup with a fixed diffusion length (determined, for example, by electrode
rotation) that corresponds to the electrode separation in the thin-layer cell. The limiting
current density for the anodic branch of the plot is higher than that of the cathodic branch
due to the Γ concentration being typically ten times higher than the I3 concentration. The
limiting current density in the two-electrode thin-layer cell configuration is determined
by the I3 concentration. The IV characteristics of the thin-layer cell can be obtained from
Figure 10.20 by adding the anodic and cathodic overpotentials for any value of the abso­
lute current density to obtain the cell voltage as a function of current density. Examples
of the resulting thin-layer cell characteristics are illustrated in Figure 10.21.
Eq. (10.55) can be linearized for small values of η to obtain a form analogous
to Ohm's law [10.92]

(10.57)

where Rct is the charge transfer resistance. In principle, measurements of the current
voltage plot of a thin layer cell close to zero voltage can be used to obtain an estimate of
356 Dye-Sensitized Solar Cells

Fig. 10.20 Plots of current density vs. overpotential calculated for mixed diffusion/electron
transfer control as a function of the standard rate constant for electron transfer. The plots can be
used to obtain the anodic and cathodic overpotentials for any given value of the absolute current
density in a thin layer cell.

Rct. However, it is also necessary to take into account the ohmic resistance of the cell,
which appears in series with Rct and can be around 15-20 Ω for a 1 cm2 cell. In practice,
Rct for a good cathode is generally lower by up to a factor of 10 as compared to the
series resistance of the cell. An accurate determination of j0 using Eq. (10.40) is thus
difficult. Impedance spectroscopy is a much better approach since it allows deconvolu-
tion of the series and charge transfer resistances (see the following section).
Thin-layer cells are also useful for characterizing blocking layers for DSCs
[10.15]. In this case, the cells consist of a platinized electrode and the electrode with
the blocking layer. Generally, the currents measured in these cells are sufficiently
small that the platinum electrode remains close to the equilibrium I3 IY potential.
Figure 10.22 contrasts the IV characteristics of uncoated fluorine-doped tin oxide
glass with those of FTO coated with a thin compact layer Ti0 2 deposited by spray
pyrolysis. It can be seen that the blocking layer provides a diode characteristic, almost
entirely blocking oxidation of I", whereas the reduction of I3 is suppressed by one
order of magnitude.
The blocking layer is clearly not ideal, since it still allows current to flow under
forward bias. The reason for this is that the Ti0 2 layer is quite highly n-type-doped
(1017-1018 cm -3 ). However, in practical terms, all that is required is that the current
density due to the reduction of the redox species should be two orders of magnitude
Characterization and modeling of dye-sensitized solar cells 357

Fig. 10.21 IV plots for a two-electrode thin layer cell, constructed based on the data shown
in Figure 10.8, showing the voltage losses associated with electron transfer and diffusion
limitations.

lower than the short circuit current density of the DSC. If this condition is satisfied,
shunting via the substrate becomes negligible. Evaluations of the properties of block­
ing layers are particularly important when the I3 IV electrolyte is replaced by one-
electron redox systems, such as Co(III)/Co(II) complexes [10.18, 10.93], or by solid
hole-conductors, such as spiro-OMeTAO [10.6, 10.9, 10.94].

10.3.4 Small-amplitude time-resolved methods


Measuring charge transport and recombination from transient photovoltage rise
and decay times
Recently, O'Regan et al. [10.95] showed that the charge transport rate and recom­
bination can be measured at open-circuit conditions from the rise and decay times
associated with a small perturbation photovoltage transient. Variation of a bias light
intensity allows the effective electron diffusion coefficient, Dn, the effective electron
lifetime, rn, and hence the effective electron diffusion length, Ln = (Dnrn)1/2, to be
determined as functions of photo voltage and electron trap occupancy.
This is a particularly useful technique since the recombination and transport in
DSCs are strong functions of the electron concentration, and thus of the voltage. At
the maximum power point, the DSC should be at a voltage as close to Voc as possible,
in order to give a high fill factor. Thus, the study of transport and recombination
358 Dye-Sensitized Solar Cells

Fig. 10.22 IV characteristics of thin-layer cells consisting of a platinized electrode and a fluo­
rine-doped tin oxide (FTO) electrode. The plots illustrate the effect of depositing a thin compact
blocking layer of Ti02 on the FTO by spray pyrolysis.

should be made close to open-circuit conditions. Moreover, the comparison of trans­


port rates and recombination between different DSC devices (modifications of the
Ti0 2 layer, electrolyte composition, etc.) and conditions (e.g., temperature) can only
be made if the ratio of the number of electrons in the traps and in the conduction band
is identical. For cells with similar trap state distributions, this condition is fulfilled
when the position of the quasi-Fermi level relative to the conduction band (Ec - nEF)
is set to equivalent values in the cells that are compared. This cannot be done when
measuring transport at short-circuit conditions since nEv varies across the mesoporous
film. As suggested by O'Regan et al. [10.95], the comparison of transport between
cells is straight-forward under open-circuit conditions since nEv through the film is
approximately uniform and measurable. By varying the bias light intensity in the
small perturbation photovoltage rise time measurement, Voc is varied and can be cho­
sen for each cell to equalize Ec - nEv between cells.
The method developed by O'Regan et al. [10.95] is also useful for three other
reasons: (i) Photocurrent transients measured at Voc need to take into account the RC
constant of the DSC device [10.50]. The rise time of the photovoltage is not limited
by the RC discharge constant, (ii) The transport rate at Voc may be slower than the
recombination rate and electrons may be lost during photocurrent transient measure­
ments at Voc. The transport rate calculated from the difference between the photocur­
rent and photo voltage decay will be very uncertain if the two rates are similar, (iii) If
Characterization and modeling of dye-sensitized solar cells 359

Ln is smaller than the film thickness, the IPCE technique described in section 10.3.1 is
useful. If on the other hand Ln is comparable or exceeds the film thickness, the IPCE
method becomes unreliable.
The photovoltage risetime measurement relies on the fact that, although no
current flows in the external circuit, charging of the substrate capacitance following
a light pulse at open circuit requires a finite time determined by diffusion of elec­
trons to the back contact from the bulk of the mesoporous oxide film. The electron
diffusion coefficient, Dn, is obtained from the time required for excess photoinjected
electrons in the mesoporous oxide film to charge the substrate capacitiance, Csub.
This time, rrise, is effectively the RC time constant corresponding to the charging of
the substrate capacitance by current flowing from the mesoporous oxide film; the
so-called chemical capacitance, Coxide, (corresponding to trapped electrons) through
the transport resistance of the oxide, Rtram> into the substrate capacitance. By analyz­
ing an electronic circuit diagram, O'Regan et al. showed that the rise time can be
expressed as:
(10.58)

If the substrate resistance is smaller than the transport resistance, the transport across
the mesoporous oxide film is the limiting factor and the transport time is given by:
(10.59)

The transport time can then be calculated from the photovoltage risetime by rearrang­
ing Eqs. (10.58) and (10.59).

(10.60)

We note that the relation between the risetime and the transport time does not involve
any assumptions regarding the physically correct model for transport.
The transport time can be related to the effective diffusion coefficient, Dn, and
film thickness, d, through:

(10.61)

The numerical factor ζ, which takes into account the distribution of the stored
electron charge throughout the film, depends on the penetration depth and direction
of the illumination [10.24]: for homogeneous illumination, it is 2.54. It follows from
Eqs. (10.50) and (10.51) that Dn can be determined from the photovoltage risetime,
provided that Csub and Coxide are known.
The capacitances in Eq. (10.50) can be determined by impedance measure­
ments, as discussed above. In their study, O'Regan et al. [10.95] calculated the capaci­
tance at each Voc and used:

(10.62)
360 Dye-Sensitized Solar Cells

where AQp is the number of electrons injected by the pulse, and AVmax is the peak
height of the transient photovoltage. Δ β ρ was determined by integrating the short-
circuit photocurrent transient caused by an identical pulse as used in the photovoltage
risetime measurement. In the calculation of C oxide from the total capacitance, a value
of 15 μ¥ cm - 2 [10.15] was employed for the parallel capacitance Csub. It should be
noted that the calculation of r trans did not have any adjustable parameters.
For the photovoltage rise and decay measurements, the DSC was held at open
circuit and was illuminated with a bias light. A light pulse was superimposed on the
bias light to generate a small increment in photovoltage (a few percent of the steady
state dc photo voltage), and the transients were recorded with, for instance, a digital
storage oscilloscope.
Typical photovoltage transients measured at varying open circuit potentials are
shown in Figure 10.23. In all the experiments presented in ref. [10.95], the photovolt-
age risetime, r rise , could be fitted to a single exponential for at least the last 50 %
of the rise (Fig. 10.23, inset), although non-exponential risetimes were occasionally
observed at early times. Alternatively, r rise can be taken as the time for the photovolt-
age to reach half of the peak amplitude. Electron lifetimes were obtained from the
photovoltage decays by fitting to a single exponential.
In trap-limited devices, such as DSCs, the comparison of transport rates between
different equipment or conditions is only valid when the Fermi level of the mesopo-
rous oxide is at the same energy difference from the conduction band edge. O'Regan
et al. [10.95] demonstrated by the above-described photovoltage transient technique
how to perform such comparisons, correcting for conduction band shifts using the
density of states distribution determined from the same photovoltage transients. They

Fig. 10.23 The transient photovoltage vs. log(time) for a dye-sensitized Ti0 2 cell at varying
bias light controlled Voc's, taken from ref. [10.95]. Pulse light: 660 nm dye laser. The pulse
intensity was the same for each transient. White bias light: 0.1-30 mW/cm2. Cell area: 1 cm2.
The inset transient shows photovoltage rise vs. linear time. The thick line is a single-exponen­
tial fit. The dye was N719 and the electrolyte was R-150, purchased from Solaronix [10.96].
Next Page

Characterization and modeling of dye-sensitized solar cells 361

also showed that the relationship between the measured transport rate and measured
charge density was consistent with the trap limited transport model.
In a recent paper [10.97], Dunn and Peter described an extension of the pho­
tovoltage transient method to high photovoltages and correspondingly short times. It
was observed that, in addition to transport-controlled substrate charging during the
photo voltage risetime, the direct injection of electrons into the substrate from dyes
adsorbed at the back contact interface occurred, and had to be corrected for, at high
dc photo voltages. The results were compared with values of Dn, r n and L n obtained by
IMPS and IMVS. The comparison showed that the two methods led to similar results
when the difference in Fermi level under open circuit and short circuit conditions was
taken into account. Figure 10.24 displays diffusion length data using pairs of D n and
r n values determined by the photovoltage pulse [10.97]. The figure also presents the
essentially constant value of L n (40 μιη) that is obtained by the regression lines in the
IMPS/IMVS analysis and takes into account that these measurements should be made
for the same quasi-Fermi level. The IMPS/IMVS data extend to photovoltages < 0.4 V
due to the quasi-Fermi level being lower under short circuit conditions.
The two methods were found to provide consistent results for photovoltages
below 0.5 V. However, within the range 0.5 to 0.7 V, the diffusion length obtained with
the photovoltage transient method was not constant and peaked at 0.6 V. The reason for
this is at present unclear but has also been observed in, for instance, ref. [10.50]. The
diffusion lengths of 40-70 μιη are at least one order of magnitude greater than the film
thickness, indicating that the electron collection efficiency was close to 100 %.
Because of the non-linear response of electron transport and recombination
as a function of the light intensity, the small-perturbation technique is a very useful

Fig. 10.24 A comparison of electron diffusion lengths calculated from linear regression fits to
IMPS/IMVS (broken line) with individual values obtained using Dn and r n pairs determined by
the photovoltage risetime/decay method (points). The DSC consisted of a 4-μιη thick mesopo-
rous Ti0 2 , a Ti0 2 blocking layer on top of the TCO contact, the N719 dye and the Γ/Ι" redox
couple in a 85:15 acetonitrile:valeronitrile solvent.
CHAPTER 11

DYNAMICS OF INTERFACIAL AND SURFACE


ELECTRON TRANSFER PROCESSES
Jacques-E. Moser

11.1 INTRODUCTION

In nanodispersed semiconductors, no significant space charge layer can be established


within particles whose dimensions are inferior to the Debye length. Rather than rely­
ing on an electric field, sustained light-induced charge separation in bulk heterojunc-
tion photovoltaic devices is based on the kinetic competition between energy and
electron transfer and charge transport processes.
In dye-sensitized solar cells (DSCs), ultrafast electron injection from a molecu­
lar excited state into the conduction band of a wide-bandgap semiconductor is key to
the initial interfacial light-induced charge separation, as it has to compete efficiently
with fast radiative and nonradiative deactivation pathways and quenching reactions.
Subsequently, dye cations produced by charge injection have to be intercepted prior
to their recombination with conduction band electrons. This charge transfer between
the oxidized sensitizer at the surface and the hole transporting medium defines to a
great extent the photon-to-current conversion efficiency of the solar cell. Finally, per­
colation of electrons between semiconductor nanoparticles, as well as hole transport
within the pore network to the cathode, has to take place with sufficient speed so as to
prevent indirect electron-hole recombination.
Figure 11.1 presents a schematic of the relevant photophysical processes and
electron transfer steps involving a dye-sensitizer molecule adsorbed on the surface of
a wide-bandgap semiconductor (S | SC) in the presence of a redox mediator (D).

Photoexcitation (11.1)
Electron injection (11.2)
Excited state deactivation (11.3)
404 Dye-Sensitized Solar Cells

Fig. 11.1 An energetic scheme of electron-transfer processes taking place after charge injec­
tion from a molecular electronic excited state S* to the conduction band (cb) of a semicon­
ductor (SC). The dashed arrows represent electron-hole recombination reactions that counter
sustained charge separation.

Back electron transfer (11.4)


Dye regeneration (11.5)
Indirect charge recombination (11.6)
The yield, η^ of a first-order kinetic reaction occurring in parallel with a second
reactive pathway is given by the simple relationship:

(11.7)

where £a and kh are the respective first-order rate constants of both parallel processes.
Hence, a reaction a will take place almost quantitatively (r/a > 0.99) provided that its
rate constant is at least two orders of magnitude larger than that of a kinetically com­
petitive process b (ka > 99 x kh).
Dye-sensitizers, which do not undergo significant intersystem crossing, are
characterized by excited state lifetimes typically in the range of r = 10-100 ps. Under
such conditions, efficient charge injection should clearly take place within a sub-
picosecond time frame. Triplet MLCT excited states of Ru(II) polypyridyl complex
sensitizers are usually much longer-lived, with r = 10-100 ns. In the presence of con­
centrated electrolytes, however, fast quenching reactions often occur in the ps time
scale. Femtosecond electron injection is therefore generally required to ensure effi­
cient initial light-induced interfacial charge separation.
The rate of the electron recapture, which takes place between the solid and
the oxidized dye species S + (Eq. 11.4), is usually observed to be slower by several
orders of magnitude as compared to charge injection rates of efficient sensitizers. In
Dynamics of interfacial and surface electron transfer processes 405

the N719 | T i 0 2 system, this back electron transfer process typically occurs in the
hundreds of μ$ to ms time scale, and such a slow charge recombination process can
be intercepted by the reaction of a reducing mediator with the oxidized dye (Eq. 11.5).
DSCs based on the sensitization of mesoporous titanium dioxide by Ru(II) complex
dyes in conjunction with the Ι3/Γ redox couple as a mediator have proved very effi­
cient when it comes to exploiting this principle. However, oxidation of I" to I 2 or I3 is
a two-electron redox process, which is intrinsically slow. Relatively high concentra­
tions of iodide are thus necessary.
Figure 11.2 shows, as an example, the temporal evolution of the oxidized state
S + of the complex dye sensitizer ds-[Ru n (dcbpyH) 2 (NCS) 2 ] 2 - (N-719). The S + spe­
cies is initially formed during photoinduced electron injection and later decays due to
reduction by a mediator or charge recombination. In this particular case, the kinetic
competition between processes (d) and (e) (Eqs. 11.4, 11.5) is minimal and results in
the formation of a long-lived charge-separated state ( e~b ... D + ) with a quantum yield
close to unity.
A thorough understanding of the functioning of successful DSCs, and perhaps
even more importantly of the details of the mechanisms resulting in a bad perform­
ance, can be gained by the study of the dynamics of individual surface electron trans­
fer reactions and charge transport processes. The knowledge acquired in performing

Fig. 11.2 Transient absorbance signals recorded upon pulsed laser excitation of N-719 | Ti0 2 .
Optical signals reflect the appearance and decay of the oxidized state S+ of the dye. The data
points at the shorter time scale correspond to the electron injection process and concomitant
formation of the S+ species (Eq. 11.2). The decay curve at a shorter time scale was obtained in
the presence of a liquid electrolyte containing 0.8 M Γ and is indicative of the dye regeneration
reaction (Eq. 11.5). The decay curve at a longer time scale is due to the back electron transfer
(Eq. 11.4) and was recorded in a pure redox-inactive solvent. Ultrafast transients were meas­
ured at a probe wavelength of 860 nm, following pumping at 535 nm. The ns-/xs data were
obtained at λ = 680 nm upon 600-nm pulsed laser excitation [11.1].
406 Dye-Sensitized Solar Cells

time-resolved analysis of such phenomena is directly relevant for the improvement of


existing systems, as well as for the design of new strategies to achieve higher energy
conversion efficiencies and a superior stability of devices. Parameters influencing the
kinetics of the electron injection, charge recombination and dye-cation interception
reactions and their interplay have recently been reviewed [11.2].

11.2 ENERGETICS OF CHARGE TRANSFER REACTIONS

The knowledge of the relative positions of the energy level of reactants is essential
for understanding the electron transfer dynamics. The energy gap between the inter­
acting levels represents the nuclear reorganization barrier to attaining a condition of
resonance. This thermodynamic aspect, along with nuclear relaxation dynamics and
electron coupling, define the kinetics of charge transfer reactions.
Surface redox processes can occur with either adsorbed or dissolved reactants.
Since electron transfer with solution species cannot be faster than what is allowed by
the diffusion, this limitation is not encountered in the adsorbed state. For excited mol­
ecules in solution, the deactivation processes are generally much faster than the aver­
age diffusion time to the surface. Hence, efficient charge injection can only take place
in the adsorbed state. Static quenching of the sensitizer and re-reduction of oxidized
species also depend critically on the association of the mediator ions with the dyed
surface. Adsorption processes at the interface are therefore very important in defining
both the thermodynamics and the kinetics of key charge transfer reactions.

11.2.1 Mesoscopic metal oxide semiconductors


Adsorption of ions and molecules on the surface
Electrophoretic measurements applied to dispersions in organic solvents of titanium
dioxide scratched from sintered mesoscopic films show that nanoparticles are consist­
ently negatively charged. The zeta-potential of a naked Ti0 2 surface in pure ethanol
is measured as being typically ζ = - 45 mV, with little variation with regard to the
preparation method. Due to the oxide having been calcinated for several minutes at
450 °C, its surface is essentially dehydroxylated. The negative charge is thus probably
due to Cl" impurities originating from the TiCl4 precursor, or to Cl" and NO3 anions
incorporated in the oxide when hydrochloric or nitric acids are used for the peptiza-
tion of particles during the preparation of the Ti0 2 paste. In such conditions, elec­
trostatic interactions should oppose the approach of anions to the surface. Efficient
adsorption of anionic carboxylated dye species, such as [Run(NCS)2(dcbpyH)2]2~ (N-
719), is however observed on the surface of Ti0 2 . In this case, chemisorption clearly
overcomes the electrostatic repulsion and renders the derivatized oxide surface even
more negatively charged, with zeta-potentials finally establishing around ζ = -54 mV
at saturation.
The surfaces of oxides are characterized by the presence at the crystal boundary
of metal centers which are not coordinatively saturated by the oxygen atoms of the
lattice. Their surface concentration on Ti0 2 (anatase) nanoparticles, for instance, was
Dynamics of interfacial and surface electron transfer processes 407

measured as Γ = 7 μιηοΐ m -2 [11.3]. These surface cations can coordinate to oxygen-


containing molecules such as water or alcohols, as well as to anions. TiIV, ZrIV, Nb v or
Al m metal centers, and to a lesser extent Zn11, are indeed strong Lewis acids and show
a particular affinity for hard bases. The adsorption strength of the anions therefore
depends on their Lewis basicity, in the typical order F~ > oxoanions > Cl" > Br~ > I" >
I3 > (CF3S02)2N~. Among oxoanions, large basicity differences exist, which translate
into various affinities for the metal oxide surfaces. Hard oxoanions, i.e., with high
charge densities and low polarizabilities, bind stronger than soft oxoanions. The fol­
lowing affinity series can be observed: OH" > PO*- > R-PO3" > R-COCT > R-SO" >
CIO4 > CF3SO3. Even though the binding constants of these anions have not all been
measured, qualitative observations of substitution processes on the metal oxide sur­
face allow for the above classification. In addition, chelation processes can be invoked
to account for the particular surface affinity of several bidentate ligands, like catecho-
late, salicyate or acetylacetonate.
Except in the case where the absorbed species are prone to the formation
of aggregates, adsorption isotherms on nanocrystalline metal oxide show a clean
Langmuir behavior and can be fitted with the corresponding equation (Eq. 11.8),
where C is the concentration of the species in solution, Γ 0 its surface concentration in
a saturated monolayer and K [M_1] is the adsorption equilibrium constant.

(11.8)

In dry ethanol, the adsorption equilibrium constant, K, for the dye-sensitizer


Ru(NCS)2(dcbpyH)2(NBu4)2 (N-719) is on the order of K= 5 x 104 M"1, while the
value of Γ0 ranges between 1-7 /zmol m~2, depending on the degree of protonation of
the dye and/or of the surface. The adsorption constants and surface concentration at
full coverage show a clear solvent dependence [11.4], which can be rationalized in
terms of competition between adsorption and solvation of the solute on the one hand,
and of competition between the adsorption of the solvent and that of the solute, on the
other hand. Carboxylated xanthene dyes such as dichlorofluorescein or eosin offer in
this respect an exemplary illustration of the effect of the solvent on the mechanism of
adsorption. Figure 11.3 displays the 3D-structure of anionic eosin-Y adsorbed on the
surface of titanium dioxide. When Ti0 2 is dehydroxylated, such as after calcination at
high temperature, adsorption takes place through the coordination of surface Ti(IV)
acidic sites by the carboxylate fonction carried by the eosin anion. In aqueous acidic
media, the titanium dioxide surface is positively charged, due to the protonation of
bridge oxygen atoms and amphoteric hydroxyl end groups. In this case, adsorption
of eosin anions takes place essentially through electrostatic interactions and possibly
H-bonding. Water molecules can also form an additional solvation layer between the
charged surface and the adsorbate. Under such circumstances, the distance between
the first accessible Ti(IV) surface ions and the conjugated 7r-system of the dye mol­
ecule is considerably increased as compared to the case of eosin adsorbed onto the
dehydroxylated surface. This change in the adsorption geometry results in a marked
increase of the time constant for electron injection from photoexcited xanthene dye
408 Dye-Sensitized Solar Cells

Fig. 11.3 The structure of the anionic form of eosin-Y dye adsorbed onto the surface of dehy-
droxylated Ti02 in a dry organic solvent (a), and on that of an hydroxylated oxide surface in
acidic water (b). The tridimensional structure of the dye was obtained from a MOPAC semi-
empirical quantum mechanical calculation.

molecules into the conduction band of Ti0 2 by at least one order of magnitude, from
typically Tinj < 30 fs in case (a) [11.5] to Tinj > 300 fs in case (b) [11.6].

Energetics of the semiconductor-liquid interface


Three important thermodynamic parameters for the electrochemical behavior of a
semiconductor-liquid electrolyte heterojunction include the conduction band edge
potential, <£cb, the flat-band potential, φ&, which is equal to the conduction band edge
potential at the interface, and the Fermi energy level, EF The potential corresponding
to the Fermi level (φ¥ = -EF/F, where F is the Faraday constant) can be considered as
the electrochemical potential of electrons in the semiconductor. The conduction band
edge represents the potential at which the first metallic center (with the exception of
intraband trap states) can be reduced.
Apart from the semiconductor material, the energy of the conduction band edge
depends on the nature and composition of the electrolyte. A specific adsorption of
ions can significantly shift the flat-band potential. For metal oxide semiconductors/
aqueous solution interfaces, the protonation and deprotonation of amphoteric sites
causes a -60 mV cathodic shift in φΆ each time the solution pH is increased by one
unit [11.7, 11.8]. Values for the flat-band potential were determined for Ti0 2 [11.9-
11.11] by measuring the absorbance of the conduction band electrons as a function of
the applied potential. A Nernstian behavior was observed with φΆ [V/NHE] = -0.14
-0.06 pH in aqueous medium. More recent findings suggest that also proton interca­
lation can play an important role [11.12]. Adsorption of other potential-determining
species, like Li+ [11.11,11.13] and Mg2+, also affect <£cb> while small cations (K+, Na+)
and anions (Γ, dyes) [11.13], surfactant molecules and surface complexing agents,
Dynamics of interfacial and surface electron transfer processes 409

influence the zeta-potential of the semiconductor. On many oxide surfaces, the den­
sity of charged sites present at the isoelectric point (PZZ) is rather low. As the solid
is charged by adsorption of potential-determining ions, the environment in which the
ions find themselves changes significantly. Consequently, the Nernst equation does
not hold in general, and a prediction of the surface potential is a quite difficult and
risky exercise.
Measurements in non-aqueous solutions have revealed that φΆ is directly related
to the Br0nsted acidity of the solvent [11.11]. A distinction between protic and aprotic
solvents is therefore more appropriate than that between aqueous and nonaqueous
media. In aprotic solvents such as CH3CN, φΆ is very negative (as much as - 2 V/SCE
for Ti0 2 ) [ 11.14], but can be raised up to 0 V by adsorption of hard cations like H+, Li+
or Mg2+, able to compensate the charges trapped on the surface [11.11].
The flat-band potential is determined by ions adsorbed on the surface and by
surface states. These states correspond to surface metal ions for which, because of an
incomplete coordination, the redox potential lies within the bandgap of the oxide. As
soon as electrons from the conduction band or from the solution reduce the surface
states, they become trapped and can only escape by transfer to an oxidizing species
across the interface, by light excitation or by hopping on the surface. On single-crystal
Ti0 2 rutile, for instance, TiIV/in surface states can be found 0.8 eV below the conduction
band edge [11.15]. For nanocrystalline particles, a broad distribution of surface states
potentials is expected, corresponding to the multiplicity of lattice planes at the surface,
with a suggested preferential localization of the traps at grain boundaries [11.10].
Figure 11.4 schematizes the energetics of electron transfer for an n-doped semi­
conductor in contact with three redox couples (A+/A, B+/B, and C+/C) in solution. The
situation is depicted as a function of a cyclic potential scan, for large particles (a-f),
and for a nanocrystalline assembly (a'-c'). The scan is started at a positive potential
with all solution species in the oxidized state (a). At that potential (φ¥ > φ&), the
conduction band is lowered by the electric field, while the potential at the interface
determined by surface states is fixed. This gives rise to a positive band-bending, which
acts as a barrier for the electron transfer, of a height equal to φ^-φ?. Thus, although
the Fermi potential lies below (/>°(C+/C), the reduction of C+ does not occur rapidly.
The excursion of the electrode potential into the flat-band region (b) results in the
reduction of both B + and C+ within a small potential range. If the electrode potential is
driven negative to φ^ the electric field causes an upwards band-bending (accumula­
tion layer) and the electrode exhibits a metallic behavior for the species with φ° < φΆ,
such as the (A+/A) couple, which is easily reduced (c) and again oxidized upon
reversal of the potential scan direction (d). If the electrode potential is driven positive
just below the flat-band situation, oxidation of B is observed (e). At more positive
electrode potentials, positive band-bending develops (f). Oxidation of C is expected
from thermodynamics as φ? < (/>°(C+/C), but the activation barrier φΆ - (/>°(C+/C) pre­
vents a fast charge injection into the semiconductor. As a result, the charge on the
weak donor species C is kinetically trapped.
The above description is valid only for semiconductor particles larger than the
width of the accumulation or depletion layer. Once the particle size decreases below this
limit, band-bending progressively vanishes, and the conduction band level eventually
coincides with the flat-band potential, reflecting the fact that the electric field created
410 Dye-Sensitized Solar Cells

Fig. 11.4 A schematic representation of the interfacial electron transfer processes between the
conduction band of a semiconductor and electroactive species on the surface, during the cyclic
sweeping of the electrode potential (a) —» (b) —» (c) —» (d) —» (e) —» (f).

by the species present on the surface extends inside the whole particle. The total poten­
tial drop between the surface and the center of a spherical nanoparticle of radius r is
obtained by solving the Poisson-Boltzmann equation and can be expressed as:

(11.9)

where:

(11.10)

is the Debye length, which depends upon the static dielectric constant ε Χ ε 0 ο ί the
material and the carrier density n0 [ 11.16]. From Eq. 11.9 it is apparent that the electric
Dynamics of interfacial and surface electron transfer processes 411

field in semiconductor nanoparticles is usually small and/or that high dopant levels
are required to produce a significant band-bending inside the particle. A flat-band
situation indeed prevails when charge migration is negligible as compared to thermal
diffusion (Αφ < kB77e). According to Eq. 11.9, this condition implies that r <
In the case of Ti0 2 anatase particles, assuming a static dielectric constant ε = 130
and a carrier density n0 = 1017 cm-3, then LD = 30 nm and r < 73 nm. If the amount of
majority carriers depleted from a semiconductor and the particle size are too small to
develop a space charge layer, the potential difference resulting from the charge trans­
fer across the semiconductor/electrolyte interface must drop in the Helmoltz layer. As
a consequence, the position of the band edges of the particulate material is expected
to shift cathodically upon electron injection and to move back during charge recom­
bination [11.17].
This situation prevails in nanocrystalline films, where the typical particle size
is on the order of 10-20 nm. As the conduction band is flat (<£cb = φ&), it is more
appropriate to use the conduction band potential <£cb as a reference rather than φΆ.
When a positive potential (φ¥ > 0cb) is applied to the contact electrode, electrons
deplete from the material turning it into an insulator. As soon as φ¥ < <£cb, electrons
start to percolate inside the nanocrystalline film. The applied field drops in the accu­
mulation layer formed near the particle surface and in the Helmholtz layer on the
side of the electrolyte. Thus, the salient characteristics of a particulate film as com­
pared to the bulk semiconductor is the absence of an electric field across the material.
Consequently, electrons do not migrate in particle assemblies, but diffuse in a random
walk process.
Intraband trap states
Energy states lying below the conceptual bottom edge of the conduction band obvi­
ously complicate the energetic scheme described to this point for interfacial electron
transfer reactions. Most of the semiconducting metal oxide materials used in hetero­
geneous light-induced charge transfer reactions contain more or less abundant trap­
ping sites in the bulk and at the surface. The presence of point defects, like oxygen
vacancies and interstitial metal ions, are indeed hard to avoid, even in single crystal
oxides. Coordinatively unsaturated surface metal ions represent efficient electron traps.
Because of their large surface area-to-volume ratio, such surface states are quantita­
tively important in small nanocrystallites. For example, in a Ti0 2 particle with a 10-nm
radius, as much as 6 % of all Ti(IV) ions are actually exposed at the surface. In mes-
oporous metal oxide electrodes, where nanocrystalline particles are sintered together,
grain boundaries are similar to dislocations and constitute two-dimensional arrays of
oxygen vacancies. In zinc oxide, luminescence experiments showed that electron trap
sites are mainly surface states [11.18-11.20], while photo-electrochemical studies on
nanocrystalline Fe 2 0 3 hematite demonstrated that, despite the large surface-to-volume
ratio characterizing films constituted of particles in the 25-75 nm size range, capture of
carriers by bulk and/or grain boundary traps is dominant [11.21].
The application of thermoluminescence, EPR [11.22-11.24] and impedance
techniques [11.15] to single-crystal and colloidal Ti0 2 led to the conclusion that at
least eight types of traps are active in this material, with energetic levels ranging from
0.2 to 0.9 eV below the conduction band edge. In addition, the existence of extended
412 Dye-Sensitized Solar Cells

states near the mid-band gap region, at about 1.4-1.5 eV below the conduction band
edge, has been suggested and discussed in terms of Ti-OH species [11.25].
Impedance and photocurrent spectroscopy studies using mesoporous anatase
electrodes have reached similar conclusions [11.26]. Results obtained with transpar­
ent nanocrystalline films (Fig. 11.5) actually suggest that a continuum of trap levels is
present near the conduction band edge characterized by a distribution of trap depths
[11.9, 11.27, 11.28]. Such a band tail of localized states is commonly observed in dis­
ordered or amorphous materials, where it is generally more appropriate to talk about
a mobility edge rather than a conduction band edge.
The importance of trapping sites on the dynamics of the charge recombina­
tion process in dye-sensitized nanocrystalline electrodes has been established by
experiments carried out under an externally applied electrical potential [11.29-11.31].
Biasing the Fermi level of T i 0 2 positive to the flatband potential renders it possible
to control the occupancy of localized states lying below the conduction band edge of
the semiconductor and to study the influence of traps on injection and recombination
processes. The nature and the amount of species adsorbed on the surface of the solid
have also been shown to influence the population and energetics of surface states
[11.3].
Figure 11.6 illustrates how trap states present in a titanium dioxide nanocrys­
talline film can favor sustained light-induced charge separation by slowing down

Fig. 11.5 The density of electronic states near the conduction band edge of nanocrystalline
Ti0 2 anatase as a function of the applied potential. The data were derived from the meas­
urement of the absorbance at λ = 850 nm of a transparent film electrode under potentiostatic
control, with the assumption that all accumulated electrons were characterized by a decadic
molar extinction coefficient ε = 106 mol-1 cm2. The Ti0 2 film was incorporated as a working
electrode in a three-electrode spectro-electrochemical cell. 10~2 M tert-butylammonium triflate
in anhydrous propylene carbonate was used as the electrolyte and the solution was thoroughly
degassed by repeated freeze-pump-thaw cycles [11.27].
Dynamics of interfacial and surface electron transfer processes 413

Fig. 11.6 The effect of the electrochemical potential bias applied to a nanocrystalline Ti02 pho­
toanode sensitized by Run(dcbpy)3+ complex dye on the-flrst order rate constant for the recom­
bination of conduction band electrons with the oxidized sensitizer species Rum(dcbpy)3+.

recombination following initial electron injection from a dye-sensitizer [11.29]. First-


order rate constants, kh, characterizing the back electron transfer from the conduction
band to the oxidized state of a Ru(II) polypyridyl complex (Eq. 11.4) were obtained
from time-resolved laser photolysis studies. When an anodic bias was applied to the
Ti0 2 electrode (φ¥ » φ^), trap states present in the bulk of the solid and at the surface
were emptied. Electrons injected in the conduction band from photoexcited sensitizer
molecules become thermalized to the bottom of the band and eventually fall into these
empty traps. Transfer from these localized states to electron acceptors at the surface is
hindered by the barrier represented by the energy requested to excite trapped carriers
up to the conduction band level. As a result, the recombination of injected electrons
with the oxidized state of the dye is a rather slow process. Polarizing the Ti0 2 film
from + 0.6 V/SCE to -0.8 V/SCE, where φ¥ = φ&, progressively fills trap states lying
below the conduction band. Photoinjected electrons can no longer be trapped and thus
recombine much faster with dye cations at the surface. The increase of the value of
the rate constant, kh, by a factor of 1000 over the same voltage range is observed, and
this increase steepers in the vicinity of the flatband potential. A further lowering of the
bias potential eventually results in an accumulation situation, where φ¥ < φ&, turning
the Ti0 2 film into a cathode.
Since surface electron traps arise mainly from incomplete coordination of metal
ions present at the interface, a coordination of these sites by adsorbates having a
Lewis base character tends to raise the energy of the surface states. If the electron
density of the ligand is strong enough, the electronic levels of the intraband states
can eventually be raised into the conduction band, implying that their trapping action
will vanish. Such an effect is likely to occur on the surface of dye-sensitized titanium
414 Dye-Sensitized Solar Cells

dioxide. Dehydroxylated surface Ti(IV) indeed constitutes electron traps, with energy
levels approximately 200 meV lower than the conduction band bottom edge. Upon
adsorption of carboxylated sensitizers onto these acidic sites, the coordination of Ti4+
surface ions by the dye's anchor provides an electron density sufficient for raising the
trap energy back to the conduction band edge level.
Rate constants for electron transfer from colloidal Ti0 2 particles to dimeth-
ylviologen in aqueous solution have been found to be strongly affected by surface
states [11.3]; the presence of traps causing the rate to decrease considerably. The
pseudofirst-order rate constant was 30 s_1 with bare Ti0 2 at pH = 3.6, but jumped to
7 x 105 s_1 with adsorbed salicylate at pH = 6, due to a sweeping away of the traps
into the conduction band by coordination of the adsorbate [11.3]. In nanocrystalline
electrodes at φ¥ < (/>cb, with species attached by a carboxylate or a phosphonate group,
the surface has less traps. Nevertheless, trap sites located in the bulk of the solid, due
in particular to oxygen vacancies |O|+/|O|0 and |0|^/|0| + , are not affected by the deri-
vatization of the surface and may still play an important role in affecting the kinetics
of charge transfer from the conduction band to an acceptor across the interface.

11.2.2 Dye sensitizer

Numerous studies based on the observed bulk photoelectrochemical effects and on


direct optical probing of the processes occurring at the solid surface have provided
evidence that the sensitizing mechanism involves, as a primary step, electron or hole
injection by the electronically excited sensitizer molecule into the semiconductor.
Alternatively, charge injection can concern the reductive or oxidative quenching of
the dye-excited state by a redox active species (a sup er sensitizer), followed by ther­
mal interfacial electron transfer:
(11.11)
(11.12)
In the case of the injection of an electron from the excited state of a molecu­
lar sensitizer into the conduction band of a semiconductor (Eq. 11.2), the photore-
dox reaction thermodynamics require the oxidation potential of the dye-excited
state (/>°(S+/S*) to be more negative than the flatband potential of the semiconductor
(φΆ = (/>cb for minute-size particles), and thus:
(11.13)
where </>°(S+/S) is the oxidation potential of the dye, ΔΕ0,ο its excitation energy, </>cb (SC)
the conduction band edge potential of the semiconductor, and F the Faraday constant.
Provided that the electronic coupling between the excited state of a dye sen­
sitizer and localized acceptor states at the surface of the solid is sufficient, direct
electron injection into trap sites is possible. This process was investigated in particular
with alizarin dye anchored to the insulating substrate Zr0 2 . Since the conduction band
edge of zirconia lies approximately 1 eV above the singlet excited state of alizarin,
electron injection into this band was not thermodynamically feasible. Nonetheless,
spectroscopic investigations have shown that, on a femtosecond time scale, the
Dynamics of interfacial and surface electron transfer processes 415

formation of the alizarin cation can be observed as a result of the electron injection into
surface trap states [11.32]. The ultrafast injection dynamics into the traps observed in
this system underline the importance of surface states for the initial charge separation
also for materials with a lower band edge, such as titanium dioxide.
The redox potential of the dye can shift upon adsorption from solution due to
Coulombic or stronger covalent interactions with the solid substrate. This potential
change can amount to several hundreds of millivolts. Since the oxidation potential
and excitation energy of the fully protonated form of a c/4Run(dcbpyH2)2(NCS)2]
complex dye (N-3) in solution are (£°(S+/S) =+1.10 V/SHE and Δ£Ό,0= 1.65 eV,
respectively, the oxidation potential of the MLCT excited state of the sensitizer is
established at (£°(S+/S*) = -0.55 V/SHE [11.33]. The flatband potential of Ti0 2 in
dry aprotic solvents can be as negative as φΆ = -1.25 V/SHE [11.31, 11.34]. In such
conditions, the conduction band of the solid would in principle be out of reach for the
dye-excited state and only deep localized sub-bandgap states could potentially act
as acceptor levels in the injection process. Surface protonation via adsorption of the
carboxylic groups results, however, in a positive shift of the flatband potential that can
amount to several hundreds of millivolts. Moreover, a complete deprotonation of the
four carboxylic groups of N-3 was demonstrated to cause its oxidation potential to
shift negatively by ca 300 mV [11.35]. Both effects combined together with the pres­
ence of traces of H 2 0 render the interfacial electron transfer from the dye's excited
state into Ti0 2 thermodynamically favorable.
Figure 11.7 illustrates the potential bias dependence on the injection quantum
yield from photo-excited Ru(dcbpy)3+ into a Ti0 2 nanocrystalline film. The lumines-

Fig. 11.7 The effect of the potential bias on the apparent quantum yield of the electron injec­
tion from the excited Run(dcbpy)3 to the conduction band of Ti0 2 [11.29]. The insert displays
the luminescence spectra of the dye adsorbed on the surface of the mesoporous semiconducting
film at two potentials. The excitation wavelength was λ = 500 nm. The oxide electrode was
immersed in aqueous 0.2 M LiC104 electrolyte at pH 3.
416 Dye-Sensitized Solar Cells

cence of the surface-adsorbed dye was found to be strongly affected by the bias volt­
age applied to the oxide electrode [11.29]. At +0.2 V, practically no emission could
be detected, due to the oxidative quenching of the excited state by charge injection.
A cathodical biasing of the potential gradually turned on the typical luminescence of
RunL3, while the quantum yield of charge injection was observed to drop from 1 to
0.5. This effect was ascribed to the filling of the electronic states of the conduction
band, leading to the increase of the energy of the lowest unoccupied levels. Biasing of
the Ti0 2 film can also result in the desorption of the sensitizer from the surface and to
a decreased injection yield [11.36].

11.3 KINETICS OF INTERFACIAL ELECTRON TRANSFER

11.3.1 Charge injection dynamics


The standard theoretical treatment for molecular electron transfer (ET) was pro­
vided by Marcus, supplying a correlation of the difference in Gibb's free energy AG
between donor and acceptor, the curvature of the potential surfaces and the reorgani­
zation energy. Within the framework of the Marcus theory, ET rates for a number of
systems could be predicted quite well and further quantum mechanical extensions
were even able to increase the area of validity. All of these electron transfer theories
are based on the assumption that the vibronic coupling is strong and that ET is con­
trolled by nuclear motion. Tunneling between the electron donor and the acceptor
states occurs when both electronic levels are made resonant by energy fluctuations
caused by the surrounding thermal bath. The situation can be illustrated by an energy
scheme (Fig. 11.8, left), displaying the situation for an ET-reaction between a donor
(D) and an acceptor species (A). In a classical view, the system can propagate along
a generalized reaction coordinate on the potential energy surface (PES) of the elec­
tronic configuration, approximated as one-dimensional parabola. Energy conserva­
tion allows an electronic transition between the initial encounter complex D-A (/)
and the final charge-separated state D+A~ (f) only at the intersection of the two PESs,
where the electron can be transferred from D to A. In the configuration where donor
and acceptor levels have identical energies, exactly at the crossing point of the / and
the/parabolas, electron transfer can occur with a certain probability, depending on
the electronic coupling matrix element between the initial state D-A and the final
charge-separated state D+ A~. Based on this microscopic view of electron transfer,
macroscopic rates can be calculated by assuming a thermal occupation of the energy
eigenstates of the D-A potential, the overall electron transfer process being mediated
by molecular vibrations [11.37, 11.38].
A fundamentally different situation for electron transfer is found in dye/semi­
conductor systems under illumination. Figure 11.8 (right) depicts a schematic of such
a system. The acceptor level is in this case the energetically broad conduction band
of the semiconductor (SC). As a consequence, the final charge-separated state energy
surface splits up into a manifold of acceptor parabolas and each point of the reactant
parabola is a transition state. In this wide-band limit case, there is no need for an
Dynamics of interfacial and surface electron transfer processes 417

Fig. 11.8 The energetic situation for a typical vibration-mediated electron transfer between a
donor (D) and an acceptor (A) described by the Marcus theory (left) and the situation prevailing
in the case of a continuum of acceptor states found in the dye-sensitization of a semiconductor
(right).

energy-matching mechanism via molecular vibrations and the rate constant for inter-
facial ET is essentially independent of nuclear factors. The only remaining parameter
determining the injection rate is the electronic coupling, representing the electronic
overlap integral between donor and acceptor states [11.37, 11.38].
The simplest kinetic model describing the charge injection as a non-adiabatic
radiationless process is derived from Fermi's golden rule. The rate constant for the
reaction can then be expressed as the product of a Franck-Condon weighted den­
sity of states (FCWD), which depends on the driving force, AG°, as well as on the
nuclear reorganization energy, Λ, accompanying the electron transfer, and on an
electronic factor that is proportional to the electronic coupling matrix element \H\
squared:

(11.14)

The FCWD is the integrated overlap of reactant and product nuclear wavefunctions
of equal energy. For a large number of accessible acceptor levels, the summation over all
the terms of the FCWD factor reduces to a pure density of final electronic states [11.37].
The rate constant &inj of the charge injection process can therefore be expressed as:

(11.15)
418 Dye-Sensitized Solar Cells

Here, £inj is the first-order rate constant for interfacial charge transfer from a sin­
gle reactant level to a continuum of electronic product states with density of states, p,
and constant electronic coupling, \H\, to all product states. The actual density of final
states, p, can be approximated by the reciprocal energy level spacing, l/hco, of the dye
cation oscillator of frequency ω, multiplied by a factor 0 < na < 1 accounting for the
density of empty electronic states available in the solid. Above the flatband energy
level, the density of acceptor states in the conduction band of a semiconductor is usu­
ally very large and the density of final states, p, is solely determined by the density of
energy levels of the dye cation (na = 1). Below the band edge, empty trap states are
present, for which the density decreases gradually at lower energies (na —» 0).
The density of accepting states, Nc (E), in the conduction band of a semicon­
ductor at the energy level E is given by Eq. 11.16 [11.39]:

(11.16)

where Ech is the energy of the conduction band bottom edge, and m*d& is the density
of-state effective mass for electrons. The latter parameter depends strongly on the
material. In Ti0 2 , for instance, m*de =πζ> 6me (where me is the electron rest mass
and the calculated density of states at least two orders of magnitude larger than that
in ZnO, in which m*dQ = 0.24 me [11.40, 11.41]. The density of states is also expecte
to be dependant upon the size of the semiconductor nanocrystallites. Indeed, a strong
quantum confinement results in widely spaced electronic levels and therefore in a
very low density of states. Although this size quantization effect would be negligible
for Ti0 2 , for which the exciton binding energy is very small, it is expected to play a
significant role for ZnO particles with diameters smaller than -10 nm.
Equation 11.15 can only be used when the electron transfer process takes place
from a single prepared excited state of the sensitizer. In the general case, the absorption
of photons, for which the energy, hv, is larger than the electronic excitation energy,
ΛΕο,ο, of the dye, leads to the population of higher vibronic levels of the molecule.
Relaxation of these vibrationally excited intramolecular states and of the whole sys­
tem along the classical reaction coordinate is expected to compete with the electron
transfer process. Under such conditions, the electronic coupling \H\ between the donor
and the acceptor states becomes a time- and excitation wavelength-dependant func­
tion and can thus no longer be readily accessed [11.38]. Figure 11.9 illustrates the
competition between fast electron transfer from an excited vibronic level of the initial
(/) state to the final (/) state and relaxation to the lower vibrational level of the reactant.
In principle, each crossing point at which ET occurs corresponds to a different value
of the electronic coupling matrix element \H\. If a normal Marcus region situation
prevails (-AG®t < Λ), vibrational relaxation of the electronic excitation state results
in an increased activation energy for the reaction. Hence, charge injection from hot
vibronic states of a sensitizer is usually kinetically more favorable than ET from the
lower vibrational level of the reactant state.
Two limiting cases can however be considered, enabling us to simply treat the
interfacial electron transfer process as though it involved a single prepared excited
state of the sensitizer: (a) Charge injection is slow enough compared to the vibrational
Dynamics of interfacial and surface electron transfer processes 419

Fig. 11.9 An illustration of the competition between fast electron transfer from hot vibronic
levels of a sensitizer and its vibrational relaxation. ET from the initial vibronic state prepared
by absorption of radiation energy hv' has a smaller activation barrier than that excited by energy
photons hv or obtained after vibrational relaxation to the lower level of the reactant state.

relaxation of the dye-excited state. In this event, electron transfer would be able to
take place only from the lowest excited state, and the injection quantum yield would
be simply controlled by the kinetic competition between the electron injection and
the decay of the excited state, (b) Charge injection is fast compared to nuclear relaxa­
tion of the excited state. In this case, interfacial charge transfer would occur from the
prepared hot vibronic level and the quantum yield for the primary injection process
would be close to unity. For both limiting cases, Eq. 11.15 would be relevant, pro­
vided that the electron transfer is non-adiabatic.
When the electronic coupling of the donor and acceptor becomes sufficiently
large (typically \H\ > 150 cm -1 = 0.7 k B r), the electron transfer is increasingly adi-
abatic and, in the absence of solvent dynamics control, the rate constant eventually
becomes proportional to a nuclear vibration frequency ωη. Under such circumstance,
the electronic coupling element does not enter into the rate expression. For other
cases, where \H\ is small enough, the value of the coupling element is required to
obtain a quantitative description of the electron transfer rate. There is obviously a
considerable interest in the role of the electronic coupling factor, as the separation
distance and anchoring geometry of the sensitizer on the surface determine its mag­
nitude. The Gamov expression (Eq. 11.17) is used to estimate the changes in \H\ in
photoinduced electron transfer, where the electron donor and acceptor are separated
by a fixed distance r:

(11.17)
420 Dye-Sensitized Solar Cells

The damping factor ß is an exponential coefficient for the decay of the electronic
wavefunction and typically has values ranging from 0.2 to 2.5 A"1. Provided that other
factors do not influence the electron transfer rate, Eq. (11.18) can be employed to
estimate the rate at a known separation distance:

(11.18)

Other parameters, such as spin changes, symmetry factors, and the relative ori­
entation of both reactants may obviously also influence the magnitude of the elec­
tronic coupling factor \H\2.

Determining injection rates experimentally


Most of the prior knowledge on bulk semiconductor-electrolyte interfacial charge
transfer is derived from steady-state photocurrent measurements achieved in photo-
electrochemical cells. Obtaining electron transfer rate constants from such an indirect
method, however, is difficult since the photocurrent depends on several other interfa-
cial and bulk processes. The rapid dynamics of electron injection can be investigated
by applying transient laser spectroscopy to colloidal dispersions or nanocrystalline
semiconductor films. Such materials are particularly amenable to time-resolved opti­
cal studies, as they display a decent transparency throughout all the visible and NIR
spectral domains. Moreover, they are characterized by a large surface area exposed
to the solution, yielding high sensitizer absorbance for only monolayer dye coverage.
Most presently studied oxide semiconductor systems, namely Ti0 2 , Sn0 2 and ZnO,
are of particular interest for the development of artificial photosynthetic and photo­
voltaic devices.
Earlier studies on dye-sensitized titanium dioxide have reported nanosecond
time constants for the injection kinetics in aqueous medium [11.42-11.44]. These
results were indirectly obtained from the measurement of the injection quantum yield
and they implicitly assumed that the interfacial electron transfer reaction competed
only with the decay of the dye-excited state. More recent studies have been based on
the same assumption, but have utilized measurements of the dye emission lifetime,
providing ns to fs time resolutions [11.45-11.48].
Direct time-resolved observation of the build-up of the optical absorption due
to the oxidized dye species S+ has been employed in a majority of recent studies. This
appears as a more reliable way of monitoring the charge injection process, as it does
not require any initial assumption on the sensitizing mechanism and is not sensitive
to the self-quenching of the dye-excited state. A powerful approach is constituted by
the direct detection by ultrafast mid-IR spectroscopy of injected electrons in the con­
duction band and in sub-band states. Using the same technique, vibrational spectra of
transient molecular species derived from the adsorbed sensitizer can in principle also
be recorded [11.41].
Various molecular sensitizers have been investigated in conjunction with metal
oxides. Organic dyes, such as xanthenes, phosphonated perylene and carboxylated
anthracene, coumarin-343, porphyrins, anthraquinones and natural anthocyanines,
have been employed as model systems. A majority of the studies on photo-induced
charge injection dynamics relevant to efficient DSCs have, however, used ruthenium(II)
Dynamics of interfacial and surface electron transfer processes 421

polypyridyl complexes, and more particularly [c/1s,RuII(dcbpyH2)2(NCS)2] (N-3), or


its partially deprotonated form [c/1s,Run(dcbpyH)2NCS)2]2~ (N-719). The choice of
the latter compound has principally been motivated by its success as a very efficient
dye-sensitizer in molecular photovoltaic cells. Upon irradiation with visible light,
adsorbed N-3 and N-719 dyes have been found to inject electrons into Ti0 2 nanocrys-
talline electrodes with a quantum yield approaching unity [11.33]. Newly developed
push-pull organic dye sensitizers recently displayed remarkable spectral properties
and stability, rendering them credible competitors to Ru(II) polypyridyl complexes.
These compounds are typically equipped with a cyanoacrylate group, serving simul­
taneously as an acceptor carrying the LUMO of the molecule and as an anchor to
the oxide surface, a thiophene bridge and a triarylamine donor moiety [11.49-11.51].
Ultrafast laser flash photolysis has been applied to dye-sensitized transparent films,
and the results are here used as examples illustrating the dynamics of the injection
process.

Example of [Ruu(dcbpy)2(NCS)2]-sensitized titanium dioxide


Figure 11.10(a) presents the transient difference spectra obtained upon nanosecond
laser excitation of N-3 in an ethanolic solution as well as of nanocrystalline tita­
nium dioxide transparent films, onto which the sensitizer was adsorbed. The dye
was excited with Λ = 605 nm output of a laser system and the absorbance change
observed immediately after the laser excitation is plotted as a function of the detec­
tion wavelength [11.52]. Luminescence quenching and photocurrent experiments
have confirmed that Λ = 600 nm excitation of the sensitizer resulted in the formation
of the charge-separated state S+|e~b (SC). The spectrum obtained upon irradiation of
dye-sensitized Ti0 2 displays a broad absorption feature, peaking around 800 nm, for
which the half-lifetime exceeds 0.5 pis. Such a lifetime is more than one order of
magnitude longer than that of the isolated dye-excited state in solution (r = 15-50 ns)
[11.33]. The recorded spectrum is comparable to that of the one-electron oxidation
product [Rum(dcbpy)2(NCS)2]+ of the complex produced by oxidative quenching of
the excited state in an alcoholic solution containing methylviologen as an acceptor
[11.53], or generated by pulse radiolysis [11.54]. This can be readily distinguished
from the spectrum of the dye-excited state obtained in solution, whose band maxi­
mum is located at Λ = 710 nm, and these observations unambiguously demonstrate
that the transient spectral feature observed upon excitation of the sensitized semicon­
ductor cannot be assigned to an excited state of the dye, but must rather be attributed
to the charge-separated state resulting from interfacial charge injection. Under such
conditions, both an LMCT transition of the - NCS ligands to the Ru(III) metal ion
center in S+ and absorption by conduction band and/or trapped electrons contribute
to the spectrum.
Further sub-picosecond data were collected. Transient results measured for
dye-sensitized Ti0 2 films were compared to those obtained for control dye-coated
Zr0 2 films, as the high conduction band edge of the latter material should prevent
electron injection. Figure 11.10(b) shows the absorption difference spectra obtained
at a time delay of 5 ps after Λ = 605 nm pulsed excitation. The spectrum obtained for
the dye-sensitized zirconia films exhibited a maximum at 710 nm, as also observed
422 Dye-Sensitized Solar Cells

Fig. 11.10 (a) Transient absorbance spectra obtained upon ns-pulsed laser excitation of N-3
dye in ethanol (1) and adsorbed on a Ti0 2 transparent film (2). The spectra were recorded 50 ns
(la, 2a) after laser excitation (A = 605 nm, 5 ns pulse duration), (b) Transient absorbance spec­
tra recorded 6 ps after ultrafast laser excitation (A = 605 nm, 150 fs pulse duration) of N-3 dye
in ethanol (1) and a freshly sensitized Ti0 2 film (2). The insert shows the temporal behavior of
the absorbance of N-3 | Ti0 2 measured at A = 750 nm with sub-ps time resolution.

for the N-3 dye in ethanolic solution, which was therefore assigned to the dye MLCT
excited state. On the other hand, the transient spectrum recorded for sensitized T i 0 2
displayed a maximum at 800 nm, which is characteristic of the dye cation. In contrast
to the data obtained for dye-coated Z r 0 2 films, the spectra measured with sensitized
T i 0 2 exhibited a certain temporal evolution for time delays of less than 5 ps. Typical
transient absorption data at a probe wavelength of A = 750 nm is shown in the insert
of Figure 11.10. The results point at a fast - 1 0 0 fs instrument response limited signal
growth followed by a slower kinetic phase extending toward several picoseconds.
Dynamics of interfacial and surface electron transfer processes 423

A detailed multi-exponential analysis of the obtained traces revealed at least three


kinetic components with lifetimes of <100 fs (35 %), 1.3 ps (22 %) and 13 ps (43 %)
[11.34, 11.55].

Non-exponential injection dynamics


The kinetics of the electron injection from N-3 have been under active study for the
last decade [11.48, 11.53, 11.56-11.60]. Studies conducted on the same system under
somewhat different conditions present only a single <50 fs phase of electron injection
[11.41, 11.45, 11.56, 11.60]. After the seminal work of Tachibana et al. [11.53], who
reported on the charge injection taking place within <150 fs (50 %) and 1.2 ps (50 %),
the study by Benkö et al. also found the electron transfer to occur in a biphasic manner
[11.57]. A first, ultrafast component was estimated to have a rise time of 28 fs and a
second, multiexponential part occurred in the 1-50 ps time domain. Although most
other recent studies seem to confirm the presence of the slower kinetic component,
the relative contribution of the latter ranges throughout the literature from 16 to 65 %
and the time constants vary from 0.7 to 100 ps with a marked non-exponential behav­
ior. This discrepancy between reported results is illustrated by Figure 11.11. Lately,
time-resolved single photon counting has been utilized to measure the electron injec­
tion dynamics in complete DSCs employing an N-719 ruthenium bipyridyl sensitizer.

Fig. 11.11 Transient absorbance signals recorded after ultrafast excitation of N-3 dye adsorbed
on nanocrystalline Ti0 2 at probe wavelengths λ = 860 nm (D) [11.60], 1.6 μηι (■) [11.58],
760 nm (O) [11.34], and 860 nm (Δ) [11.59] as reported by four research groups. In all cases,
the temporal evolution of the signal, due to absorption by oxidized dye species S+ and/or
conduction band electrons, provides a direct measurement of the kinetics of the same charge
injection process. The insert shows the effect on the observed kinetics of the concentration of
the ethanolic dye sensitizer solution used to load the oxide surface: 0.3 mM (D) and 0.015 mM
( · ) [11.60].
424 Dye-Sensitized Solar Cells

While the quenching of ca 50 % of the dye's triplet state photoemission could not be
time-resolved, 50 % of the excited states appeared to decay with a half time of 200 ps
[11.48, 11.61].
The observed intricate injection kinetics were rationalized by Sundström et al.
in terms of a two-state mechanism; the fast and slow components being attributed
to the injection from the singlet and the triplet excited states of the ruthenium com­
plex, respectively [11.57]. Other authors have attributed the multiphasic nature of the
injection to sensitizers adsorbed on energetically different sites or in various spatial
configurations at the surface of the mesoscopic titania films [11.62]. The origin of
such complex electron transfer kinetics remains unclear. The large spread of time
constants experimentally determined for the injection process and the obvious contra­
dictions between data reported in the literature question the proposed interpretations.
The participation of various singlet electronic excited states of the Ru(II) complex to
the reaction is ruled out as one fails to observe any effect of the excitation wavelength.
Nor has a significant difference been observed in the results obtained for dyed Ti0 2
films exposed to air and in propylene carbonate, thus apparently excluding possible
effects due to the solvation dynamics. Direct interfacial electron transfer to various
localized defect states has been associated to different electronic coupling elements
and has therefore resulted in a wide distribution of rate constants. The occupancy of
these trap states can be modulated by sweeping the Fermi level below the flatband
energy upon applying an external electrical bias. Modulation of the applied potential,
does however not appear to result in any noticeable change in the injection yield and
fast kinetics [11.34]. On the other hand, adsorption of potential determining cations,
such as Li+, causing the flatband potential of the semiconductor to shift positively,
apparently affect the electron injection rate [11.34, 11.63]. These observations sug­
gest that the multiple time constants result from heterogeneities in the energetics of
the nanocrystalline Ti0 2 films.

Effect of surface dye coverage on the injection kinetics


The adsorption of dye molecules on a variety of surface sites and with various anchor­
ing geometries may also cause an intricate kinetic outcome [11.60, 11.64]. Results
obtained with eosin-sensitized aqueous Ti0 2 colloids have demonstrated that kinetic
heterogeneity could arise from intermolecular interactions and a wide distribution
of geometrical conformations of loosely bound dye molecules on the surface [11.6].
By carefully controlling the deposition of N-3, N719 and of their parent heteroleptic
[RuII(4,4/-dicarboxylic)(4,4/-dinonyl-2,2/-bipyridine)(NCS)2] (Z-907) complex onto
nanocrystalline Ti0 2 , oxide films sensitized by much less than one dye monolayer dis­
play essentially monophasic injection kinetics with a time constant of less than 20 fs
[11.60]. On the other extreme, adsorption of the sensitizer on the oxide surface for a
prolonged time or from concentrated dye solutions leads to a marked decrease of the
initial rapid phase and to the appearance of slower kinetic components. These extend
to tens of picoseconds and account for up to 26 % of the overall number of S+ species
formed during the injection process (Fig. 11.11, insert). The results show that the slow
kinetic phase observed for electron injection from N-3 or N-719 into Ti0 2 is not an
intrinsic property of the dyes as previously suggested. It is rather a consequence of
Dynamics of interfacial and surface electron transfer processes 425

aggregation and poor ordering on the semiconductor surface of dye molecules, whose
electronic coupling with acceptor levels of the semiconductor thus becomes reduced
by several orders of magnitude.
In the presence of high concentrations of iodide ([Γ] > 0.8 M) - a situation typ­
ically encountered in ionic liquids based on iodide melts where I" concentration can
reach 6 M - significant reductive quenching has been observed for aggregated sam­
ples. Up to 25 % of the dye-excited states were intercepted in these conditions before
they could inject [11.65]. The reaction of aggregated dye-excited states with iodide
appeared to take place within a few tens of ps. In addition, N-719 and Z-907 anions
formed by reductive quenching of loosely bound dye-excited states on the surface
were found to be extremely long-lived and to decay with a time constant of ca. 1 ms.
Despite the large thermodynamic driving force for charge injection from dye anions
into the semiconductor conduction band, this reaction does appear to take place, sig­
nifying that a non-negligible fraction of the potential photocurrent could be lost in
DSCs through the reductive quenching route in the presence of high iodide contents.
The mechanism of the one-electron oxidation of I" by the dye-excited state, S*, or by
the oxidized species is believed to be similar and is discussed in section 11.4.1.

Injection from hot vibronic excited states of the dye


There is now compelling evidence that the fastest kinetic phase of electron injec­
tion in c/1s,[RuII(dcbpy)2(NCS)2]-sensitized nanocrystalline titanium dioxide films
takes place in the femtosecond regime. The vibrational relaxation of the dye-excited
state, on the other hand, is expected to occur typically within 0.4-1 ps (ky = 1012s_1).
Consequently, observed injection rate constants on the order of kx = 1013-1014 s_1 cer­
tainly preclude thermalization of the dye-excited state, S*, to its lowest vibrational
level V = 0 prior to the reaction, and suggest that charge transfer can occur directly
from hot (ν' > 0) excited sensitizer molecules. To experimentally test this possibility,
systems were designed in such a way that the V = 0 energy level of the electronically
excited state of the dye lay below the bottom of the conduction band of the semicon­
ductor, where the density of available acceptor states was low (Fig. 11.12). Under
such conditions, charge injection from vibrationally relaxed excited molecules of the
sensitizer was either slow or unfeasible. However, when electron injection from a hot
vibronic state of the dye is able to compete successfully with its nuclear relaxation
(k\ >ky\ charge injection becomes possible for higher excitation photon energies,
and an excitation wavelength dependence of the quantum yield Φ, = k\l(k\ + kv) is
obtained [11.52, 11.66, 11.67].
Whether hot injection could eventually lead to a practical device system ren­
dering it possible to exceed the single-junction Shockley-Queisser limit of 31 % for
photovoltaic power conversion efficiency is debatable [11.2]. Collecting hot conduc­
tion band electrons before their thermalization and trapping seems indeed to be more
difficult to realize than hot injection from upper vibrational levels of electronically
excited molecular sensitizers.
Observations of an excitation wavelength dependence of the charge injection
process demonstrate that photoinduced interfacial electron transfer from a molecular
excited state to a continuum of acceptor levels can take place in competition with the
426 Dye-Sensitized Solar Cells

Fig. 11.12 An energy scheme for a dye sensitizer, S, adsorbed onto semiconductors character­
ized by varying values of Ecb. To the left, charge injection is thermodynamically feasible for
the dye-excited by photons of any energy hv > AE0,o- In the situation depicted to the right, only
photoexcitation to hot energy states (h^ > AE0t3) allows for the injection process.

relaxation from upper excited levels. The rather slow growth of the injection quantum
yield above the energy onset suggests that it actually reflects the density of acceptor
states in the solid that are present below the conduction band edge. In conditions where
the injection quantum yield is unity (k\ » kv) and electron transfer takes place to the
conduction band of the semiconductor (na =1), the occurrence of the electron transfer
process from a single prepared state S*(v' > 0) validates the simple model of Eq. 11.15
and makes it possible to estimate the electronic coupling matrix element \H\. Assuming
a frequency of the dye cation oscillator tö < 1500 cm-1, calculations would give a value
of \H\ > 500 cm-1 (= 2 k7) from a rate constant kx = 5 x 1013 s_1 measured typically fo
[Run(dcbpy)2(NCS)2]-sensitized nanocrystalline Ti0 2 . Although the value employed
for the cation vibration frequency, tö, and therefore that determined for \H\, are here
probably over-estimated, this figure corresponds to a rather strong electronic coupling
and indicates that the electron injection rate has reached the adiabatic limit.

Parameters influencing the rate of electron injection


The Franck-Condon weighted density of states factor, comprising parameters such as
the reaction free energy, AG°, the nuclear reorganization energy, Λ, and the tempera­
ture, T, is expected to play only a negligible role in systems that are kinetically near
optimum in terms of the Marcus theory and that are characterized by a large number
of acceptor states. According to Eq. (11.15), the rate of interfacial electron transfer
is controlled only by the electronic coupling matrix element, \H\, and the density of
the acceptor states. The activationless nature of the charge transfer process has been
experimentally confirmed by the observation of temperature-independent injection
kinetics [11.68]. Other reported data show that, under energetically favorable condi­
tions, the rate of electron injection is not controlled by the energetics of the sensitiz­
er's excited state, nor by the medium reorganization, but rather by the density and
occupancy of electronic states in the solid [11.34].
Dynamics of interfacial and surface electron transfer processes 427

The rate of electron injection has been determined in nanocrystalline ZnO and
MoS2 quantum dots. In d^Run(dcbpy)2(NCS)2]-sensitized zinc oxide, highly non-
exponential injection kinetics were measured, which could be fitted by three expo­
nential components with <1 ps (18 %), 42 ps (46 %), and 450 ps (36 %) rise times
[11.69]. The multi-exponential kinetics could be described by a model assuming a
Gaussian distribution of electronic coupling between the π* orbital of the ruthenium
complex ligands and the accepting orbitals in the solid. The observed kinetics were
approximately 10-100 times slower than those measured on Ti0 2 under similar condi­
tions. This important difference can be rationalized by the density of states in the con­
duction band of ZnO, which is estimated to be about two orders of magnitude smaller
than that in titanium dioxide [11.11]. In small MoS2 nanoclusters, an electron injec­
tion time of 250 ps has been observed upon sensitization by N-3 dye [11.64]. Ground-
state spectra and adsorption properties of the dye suggested in this case that electronic
coupling between the carboxylated ligands and the semiconductor surface could not
be very different from that on Ti0 2 and ZnO. The quite slow kinetics observed on
MoS2 were therefore likely to be the result of a far smaller density of acceptor states
in nanoclusters, exhibiting a strong quantum-confinement.
On titanium dioxide, the photoinduced charge injection process was reported
to take place on time scales ranging from less than 10 fs to several microseconds,
depending on the sensitizer and conditions used [11.70]. Such a spread of the values
of k{ over nine orders of magnitude can be accounted for only by very different values
of the electronic coupling between the dye-excited state and the acceptor orbitals
at the surface of the semiconductor. Using Eq. 11.15, and assuming for all systems
na = 1 and an average collective vibrational mode frequency of the dye oxidized
state 05 = 1500 cm-1, the electronic coupling matrix element can be calculated for
each sensitizer. Obtained values of \H\ vary from 0.02 cm -1 to > 1600 cm-1. \H\ val
ues > 200 cm -1 (=kBT) are however hardly compatible with the non-adiabatic assump­
tion of Fermi's golden rule (Eq. 11.14) and injection processes occurring in less than
50 fs should thus be considered to proceed adiabatically. Considering Eq. 11.17, and
assuming a through-space damping factor /3 = 1.2 A -1 , the full range of &,· figures
between 2 x 105 s_1 and 2 x 1013 s_1 implies a difference in the electron transfer reac­
tion distance on the order of 15 A between the slowest and the fastest system.
Various types of association of the sensitizer with the oxide surface could
explain such a difference. The strong electronic coupling prevailing for an efficient
sensitizer is generally the result of the anchoring of the dye molecule onto the semi­
conductor surface through a moiety carrying its lowest unoccupied molecular orbital
(LUMO). This situation is clearly encountered in carboxylated Ru(II) polypyridyl
complexes, coumarin or alizarin dyes. The examples provided by xanthenes and N-3
dye- sensitization of titanium dioxide demonstrate that the mode and geometry of
adsorption of sensitizers at the surface of the semiconductor can strongly affect the
ultrafast photoinduced charge injection dynamics. A decrease of the donor-accep­
tor electronic coupling is likely to occur with dye molecules loosely associated to
the charged semiconductor surface through electrostatic interaction and/or hydrogen
bonding. Results obtained for the eosin-sensitized aqueous titanium dioxide colloids
are exemplary of the sensitivity of the dynamics of interfacial electron transfer upon
surface and environmental conditions in the weak-coupling case [11.6]. A dispersion
428 Dye-Sensitized Solar Cells

of dye monomers within a nanometer-thick poly vinyl-alcohol adlayer yielded a broad


distribution of distances separating the sensitizer's excited states from the reactive
surface. In this situation, kinetic parameters for charge injection in the conduction
band of Ti0 2 were found to cover a large time span, from typically 200 fs to hundreds
of picoseconds, and were only limited at longer times by radiative and nonradiative
decay of the dye-excited states.
For Ru(II) bipyridyl sensitizers as well as organic dyes, a strong electronic
coupling between the π* molecular orbital of the dye-excited state and the empty
TiIV-3d orbital manifold of the semiconductor can be achieved by directly linking
the sensitizer's moiety carrying the LUMO to the surface. Both carboxylic and phos-
phonic anchoring groups are fairly good at attaching dye sensitizers onto the sur­
face of Ti0 2 . Nevertheless, they are not equivalent in terms of electronic coupling.
Carboxylic groups attached in the 4,4' position on bipyridyl ligands are conjugated
with the π-system of the aromatic core. As a result, the LUMO carried by the lig­
ands can extend through this conjugated bridge in close proximity to the first empty
d-orbitals of surface Ti(IV) ions and thus ensure a strong electronic coupling with the
conduction band acceptor levels. 7r-conjugation is not possible through phosphonic
groups, which then act as insulating bridges between the bipyridyl ligand and the
surface. Although the adsorption equilibrium constant is larger for phosphonated lig­
ands than for their carboxylic counterpart, phosphonated dyes are characterized by a
weaker electronic coupling for charge injection. Figure 11.13 provides an illustration
of the results obtained in the determination of the injection dynamics from two similar
Run(tpy)(NCS)3 dyes, whose terpyridyl ligand carries either a single carboxylic or a

Fig. 11.13 Ultrafast transient absorption data of RunL(NCS)31 Ti02 systems recorded with
pump and probe wavelengths of 530 nm and 560 nm, respectively. The structure of the terpy­
ridyl ligand, L, is indicated for each set of data. The absorbance change is due to the decay of
the dye-excited state during the electron injection process. The kinetics of the injection from
the carboxylated dye can be fitted by time constants 120 fs (50 %) and 800 fs (50 %). Time
constants for the phosphonated sensitizer are 800 fs (50 %) and 17 ps (50 %) [11.71].
Next Page

Dynamics of interfacial and surface electron transfer processes 429

phosphonic anchoring group. The rate constant obtained for the faster kinetic com­
ponent is k{ = 8.3 x 1012 s_1 for the carboxylated compound and hx = 1.2 x 1012 s_1
for the phosphonated sensitizer. Applying again Eq. 11.17 and assuming a distance
damping factor ß = 1.0 A"1, this 7-fold difference corresponds to an increase in the
distance for electron transfer by ca 2 A, which is roughly equivalent to the dimension
of the phosphonic group spacer.
The electronic coupling for interfacial ET can be deliberately diminished by
increasing the distance separating the LUMO of the dye from the surface of the semi­
conductor material. The dependence of the multiphasic injection dynamics upon this
parameter could serve to discriminate between the various possible sources of kinetic
heterogeneity. This can be achieved, for instance, by inserting insulating spacer units
between the chromophore and the anchoring group of the dye. Lian and co-workers
have studied the bridge-length dependence of ultrafast charge injection from rhenium-
polypyridyl complexes to Ti0 2 and Sn0 2 films and have suggested that the transition
between the strong (adiabatic) and weak coupling (non-adiabatic) cases takes place
for transfer distances increased by only one -CH 2 unit length (~3 A) [11.58, 11.72].
The lengthening of the bridge spacer does not always lead to slower kinetics, particu­
larly if the linker is too flexible or when the molecule can adopt a tilted orientation on
the surface. In an attempt to circumvent this problem, rigid oligophenylene bridges
[11.73, 11.74] and tripodal linkers oxide [11.75] were synthesized to anchor sensi-
tizers to the surface of semiconductor oxide. A sub-picosecond injection rate was
observed over a distance of more than 20 A, with apparent damping factor values on
the order of ß = 0.04 A -1 , indicating a significant delocalization of the excited state
over the rigid spacer arm [11.74, 11.75]. Perylene-based tripodal sensitizers were
studied in conjunction with other perylene sensitizers under ultra-high vacuum condi­
tions [11.76]. Injection time constants ranging from 13 fs to 4 ps were measured and
an exponential dependence upon the ET distance was observed with a damping factor
ß = l A -1 , compatible with through-space electronic tunneling.
The distance dependence was investigated by other means, using the phos­
phonated RuII(4/-P03-tpy)(NCS)3 dye adsorbed on Ti0 2 core particles coated by
an insulating A1203 shell of increasing thickness. Experimental results, displayed
in Figure 11.14, demonstrated that electron injection occured with a relatively high
quantum yield for tunneling barriers as thick as 2-3 nm. Neglecting ultrafast injec­
tion, which is likely to be due to dye molecules directly attached to Ti0 2 from holes
in the alumina layer, a biphasic injection kinetic was observed, resulting again in
ß = 0.04 - 0.11 A -1 . As the barrier to the conduction band of bulk, crystalline A1203 is
very large (>3 eV), a ß value of approximately 1 A -1 was expected. In nm-sized layers
made of amorphous aluminium oxide, empty states should however exist at a poten­
tial of -1.5 V/SHE. This potential corresponds precisely to the oxidation potential
of the excited state of the dye and therefore indicates that amorphous alumina could
directly mediate electron transfer from the dye-excited states to the conduction band
of Ti0 2 [11.73].
An interesting example of a strong electronic coupling case is provided by the
alizarin | Ti0 2 system [11.32, 11.77]. Chelation of surface Ti(IV) sites by alizarin
(1,2-dihydroxy-anthraquinone) resulted in the red-shift of the absorption spectrum
of the dye by 70 nm due to deprotonation of both hydroxy groups. The question has
CHAPTER 12

IMPEDANCE SPECTROSCOPY: A GENERAL


INTRODUCTION AND APPLICATION TO
DYE-SENSITIZED SOLAR CELLS
Juan Bisquert and Francisco Fabregat-Santiago

12.1 INTRODUCTION

Impedance Spectroscopy (IS) has become a major tool for investigating the properties
and quality of dye-sensitized solar cell (DSC) devices. This chapter provides an intro­
duction of IS interpretation methods focusing on the analysis of DSC impedance data.
It also presents a scope of the main results obtained so far. IS gives access to funda­
mental mechanisms of operation of solar cells, for which reason we discuss our views
of basic photovoltaic principles required to realize the interpretation of the experi­
mental results. The chapter summarizes some 10 years of experience of the authors
with regard to modeling, measurement and interpretation of IS applied in DSC.
A good way to start this subject is a brief recollection of how it evolved over
the first years. The original "standard" configuration of a DSC [12.1] that emerged in
the early 1990s is formed by a large internal area constituted of a nanostructured Ti0 2
semiconductor, connected to a transparent conducting oxide (TCO) and coated with
photoactive dye molecules. It is furthermore in contact with a redox 1713 electrolyte
that is in turn connected to a Pt-catalyzed counterelectrode (CE). The DSC was ini­
tially developed to be a photoelectrochemical solar cell. Electrochemical Impedance
Spectroscopy (EIS) is a traditional method, central to electrochemical science and
technology. Interfacial Electrochemistry usually investigates interfacial charge trans­
fer between a solid conductor (the working electrode, WE) and an electrolyte. This is
done with a voltage applied between the WE and CE, with the assistance of a refer­
ence electrode (RE), rendering it possible to identify the voltage drop at the interface
between the WE and the electrolyte. In addition, the electrolyte often contains a salt
that provides a large conductivity in the liquid phase and removes limitations by drift
transport in an electrical field. Electrochemistry is thus mostly concerned with interfa-
cial charge transfer events, possibly governed by diffusion of reactants or products. It
458 Dye-Sensitized Solar Cells

is with EIS possible to readily separate the interfacial capacitance and charge-transfer
resistance, as well as to identify diffusion components in the electrolyte. A good intro­
duction to such applications is given by Gabrielli [12.2].
In solid state solar cell science and technology, the most commonly applied
frequency technique is Admittance Spectroscopy (AS). By tradition, AS denominates
a special method that operates at reverse voltage and evaluates the energy levels of
majority carrier traps (in general, all those that cross the Fermi level) as well as trap
densities of states [12.3]. In work on DSCs and other solar cells, we may be interested
to probe a wide variety of conditions. Consequently, we generally use the denomina­
tion Impedance Spectroscopy (IS) when referring to the technique applied in this
context (rather than EIS or AS).
Before the advent of DSC, IS had been largely applied in photoelectrochem-
istry [12.4, 12.5]. This is a field widely explored since the 1970s, using compact
monocrystalline or polycrystalline semiconductor electrodes for sunlight energy con­
version [12.6-12.8]. In these systems, IS provides information on the electronic car­
rier concentration at the surface, via Mott-Schottky plots (i.e., the reciprocal square
capacitance versus the bias voltage) as well as on the rates of interfacial charge trans­
fer [12.9-12.11]. Several important concepts, later to be applied in DSC, where estab­
lished at that time, such as the bandedge shift by charging of the Helmholtz layer and
the crucial role of surface states in electron or hole transfer to acceptors in solution
[12.9, 12.10, 12.12-12.14]. Nonetheless, it was clearly recognized that applying IS
in these systems is not straightforward, for example due to the presence of frequency
dispersion that complicates the determination of parameters [12.15]
It was natural to apply such well-established electrochemical methods to DSC
and several groups have done so [12.16-12.19]. However, in the early studies, it was
necessary to clarify a conceptual framework of interpretation which took several years.
On the one hand, the early diffusion-recombination model [12.20] was generally
adopted for steady-state techniques and produced very good results when extended to
light-modulated frequency techniques [12.21]. In this approach, the only role of the
applied voltage is to establish the concentration of electrons at the edge of the Ti0 2 in
contact with TCO [12.20, 12.21]. On the other hand, classical photoelectrochemical
methods heavily rest on the notion of charge collection at the surface space-charge
layer, while diffusion is viewed as an auxiliary component, at best [12.22]. Thus,
in photoelectrochemistry of compact semiconductor electrodes, the main method to
describe the system behavior is an understanding of the electric potential distribution
between the bulk semiconductor and the semiconductor/electrolyte interface [12.7].
Owing to these conflicting approaches, in the DSC area there were many discus­
sions concerning the distribution of the applied voltage as internal "potential drops",
the origin of photovoltage, screening, and the role of electron-hole separation at the
space-charge region [12.23-12.27]. This is understandable since the DSC is a porous,
heterogeneous system, and in models of systems with a complex morphology, it is
generally difficult to match diffusion control with a precise statement regarding the
electrical potential distribution. The key element for progress is to adopt a macro-
homogeneous approach and focus in the spatial distribution of the Fermi level. This
method emerged in the DSC area [12.24, 12.28-12.30] and eventually led to gener­
alized photovoltaic principles based on the splitting of Fermi levels and the crucial
Impedance spectroscopy 459

role of selective contacts [12.31-12.34]. Another central concept that appeared in the
DSC area was a "conduction band capacitance" [12.26, 12.28, 12.30], later to be
generally defined as a chemical capacitance [12.35]. This capacitive element is nor­
mally absent in classical photoelectrochemistry but is key for the interpretation of
frequency-resolved techniques in DSC. Also important was the recognition [12.26,
12.36] that nanostructured Ti0 2 should be treated as a disordered material, much like
the amorphous semiconductors [12.37-12.39], with electronic traps affecting not only
the surface events, but any differential/kinetic measurements, including the chemical
capacitance [12.35], recombination lifetime and transport coefficients [12.40].
The passage from established ideas of photoelectrochemistry to those best
suited to the DSC have inevitably rendered it necessary to treat the porous-mixed
phase structure of the DSC. Electrochemistry was already evolving in this direction for
some decades, first with the description of porous electrodes [12.41], and then, with
the introduction of truly active electrodes that become modified under bias voltage,
such as intercalation metal-oxides [12.42], conducting polymers [12.43] and redox
polymers [12.44]. Especially important is the work of Chidsey and Murray [12.44],
which shows the modification of the diffusion coefficient in the solid phase, as well as
the capacitance of the solid material as a whole, in opposition to the standard interfa-
cial capacitance. In the analysis of these systems, either porous or not, the importance
of coupling transport elements with interfacial and/or recombination components for
a proper description of IS data was well recognized. Transmission line models pro­
vide a natural representation of the IS models and are widely used [12.43, 12.45].
As demonstrated in Figure 12.1, transmission line models incorporating fre­
quency dispersion, which is ubiquitous in disordered materials, have been developed
and applied to nanostructured Ti0 2 used in DSC. A very good realization of the model
was soon found in the experiment, as shown in Figure 12.2 [12.46]. Later, diffusion-
reaction models were solved for IS characterization, and the models where put in rela­
tion to both nanostructured semiconductors and bulk semiconductors for solar cells
[12.47]. Disorder was included also in generalized transmission lines for anomalous
diffusion [12.48]. In addition, the role of macroscopic contacts was analyzed in gen­
eralized transmission line models, as shown in Figure 12.1(b) [12.49], and this effect
would take relevance as a result of the TCO contribution to the measured impedance
[12.50, 12.51].
The calculation of the diffusion-recombination impedance [12.47] opened the
way for a direct measurement of conductivity of electrons in Ti0 2 by IS [12.52],
which provided a good validation of the method. Further, the diffusion-recombination
impedance also naturally reveals [12.47] the chemical capacitance of electrons in
nanostructured Ti0 2 (associated to the rise of the Fermi level), which also appears in
measurements of cyclic voltammetry (at slow scan rates) [12.53] and electron lifetime
[12.54].
Application of these IS methods and models to DSC [12.51] demonstrated that
IS provides a picture of the energetics of Ti0 2 , which is a crucial tool for compar­
ing DSC configurations [12.55]. It also showed that it was possible to simultane­
ously obtain the parameters for transport and recombination at various steady-state
conditions of a DSC, which is an unsurpassed power of the technique. The trends
of the electron diffusion coefficient [12.51] where similar to those found previously
460 Dye-Sensitized Solar Cells

Fig. 12.1 (a) A general two-channel transmission line equivalent circuit for a porous electrode
or diffusion coupled with recombination, with blocking boundary conditions at both chan­
nel ends [12.46]. (b) The two-channel transmission line with generalized boundary conditions
[12.49]. Notice that the ZA box corresponds to the electrical properties of the electrolyte/
substrate interface, although it is not drawn precisely at that point for the sake of convenience
of representation.

by L. M. Peter and coworkers by light-modulated approaches [12.56]. The electron


lifetime derived from IS measurements was also consistent [12.55] with that obtained
from open-circuit voltage decays [12.54, 12.57]. The variation of parameters with
the bias voltage (correspondent to the electron Fermi level) observed by IS and other
methods was related to multiple trapping characteristics in an exponential distribution
of states [12.33, 12.58]. This subject has been recently summarized in several review
articles [12.59-12.61].
The consistency of the various experimental methods has provided great con­
fidence in the significance of modeling and experimental tools. The usefulness of IS
for DSC characterization has become apparent, since IS renders it possible to obtain
a complete picture of the different device aspects [12.18, 12.19]. Several groups have
presented detailed and systematic IS characterizations of DSCs [12.62-12.64]. The
literature concerning the application of IS in DSC is very large and we do not aim to
cite all the contributions. Rather, we highlight a paper on high efficiency DSC [12.65]
which provides excellent examples of diffusion-recombination impedances, a full
Impedance spectroscopy 461

Fig. 12.2 Impedance Spectroscopy of a 8-μηι thickfilmof nanostructured Ti02 (10-nm ana-
tase nanoparticles) in aqueous solution at pH 2, with -0.250 V bias potential vs. Ag/AgCl in the
dark and under UV illumination. The lines arefitsto the model of a version of the transmission
line in Figure 12.1(a) [12.46].

analysis of electron transport data, as well as the reconstruction of the current density-
potential (j-V)curve from the resistance obtained by IS. Subsequently, IS has been
applied in a variety of important configurations of DSC, such as those using ionic
liquids [12.66], ordered Ti0 2 nanotubes [12.67], and solid hole conductor [12.68].

12.2 A BASIC SOLAR CELL MODEL

12.2.1 The ideal diode model


Many general aspects of solar cell operation can be understood starting with an ideal
model that represents optimal performance. Figure 12.3(a) shows the steady-state char­
acteristic j-V curve of a solar cell. This curve was drawn using the ideal diode model:

(12.1)
462 Dye-Sensitized Solar Cells

Fig. 12.3 (a) Theoretical calculation of the current density-voltage characteristic of a solar cell
(ideal diode model) with ysc = 25 mA cm-2, m = 1 and Voc = 0.8 V. Also indicated are the differ­
ent regions of the applied bias voltage and of the dominant current, as well as the calculation
of the dc resistance R^ = dj/dV at a particular point (V0J0). (b) The power output of the solar
cell. The left vertical axis is normalized to the incident power of 1 sun and gives the conversion
efficiency, and the right axis normalization gives the fill factor at the maximum point.

Here, j is the electrical current density, V is the voltage difference between the
contacts, j s c is the short-circuit current density, jd is the dark reverse current density,
q is the positive elementary electrical charge, kB is Boltzmann's constant and 7 i s the
absolute temperature. The coefficient m is an ideality factor, and the "ideal" model
corresponds to m = 1. From Eq. (12.2), we obtain the open-circuit voltage Voc:

(12.2)
Impedance spectroscopy 463

and we can also write Eq. (12.1) in terms of Voc

(12.3)

Bias voltage is denoted "forward" when it injects charge in the solar cell and
induces recombination. Otherwise it is referred to as "reverse". By changing the
illumination intensity Φ0, one can trace curves similar to that in Figure 12.3(a) with
other values of j s c and Voc. The values and shape of these curves for a given solar cell
allow us to determine the energy conversion efficiency of the photovoltaic device,
Figure 12.3(b). Another crucial parameter is the fill factor (FF), which is the maxi­
mum electrical power delivered by the cell with respect to j s c Voc, Figure 12.3(b).
A high FF requires that the current remains high at the maximum power point. This is
obtained if the j-V curve is reasonably "squared" as in Figure 12.3(a).

12.2.2 Physical origin of the diode equation for a solar cell

It is important to clarify the physical interpretation of the diode equation. We consider


a slab of p-type semiconductor with thickness L. At a position x, n is the density of
minority carriers (electrons), and Jn the flux in the positive x direction. The conserva­
tion equation can be written as:

(12.4)

where G<p is the rate of optical photogeneration (per unit volume) due to the illumi­
nation intensity Φ0 (photons-cm-2), while G d is the rate of generation in the dark by
the surrounding blackbody radiation. Un is the rate of recombination of electrons per
volume. A simple and important model is the linear form, with electron lifetime r 0

(12.5)

Eq. (12.4) must hold locally, in equilibrium, therefore, assuming Eq. (12.5),
we have

(12.6)

where n0 is the carrier density in dark equilibrium. This is due to the rate of generation
in dark equilibrium, by detailed balance principle, equilibrating the recombination
rate [12.31]. A similar constraint on Gd applies for any recombination model.
The flux of electron carriers with the diffusion coefficient D0 relates to the gra­
dient of concentration by Fick's law

(12.7)
464 Dye-Sensitized Solar Cells

While Eq. (12.4) can be solved for any kind of generation profile and bound­
ary conditions, we now adopt certain assumptions that lead to the central diode
model (12.1) in the simplest way. We assume that the photogeneration of carriers is
homogeneous, and we consider that the transport of electrons is very fast. It can thus
be assumed that D0 is extremely large, implying that the gradient of concentration
required to maintain the flux is very small. With these assumptions all the quantities in
Eq. (12.4), except the carrier flux, become independent of position. We now integrate
between 0 < x < L and obtain

(12.8)

The next condition required is to assume that the semiconductor is supple­


mented with ideal selective contacts to form a solar cell, as shown in Figure 12.4
[12.33]. Consequently, the left contact extracts all the arriving electron carriers.
The electrical current density in the positive x direction is:
(12.9)

and the right contact blocks the electrons perfectly:


(12.10)

Therefore, the output current at time t is:

(12.11)

Fig. 12.4 A basic model of a solar cell formed by a light absorber and two selective contacts for
electrons and holes. The image shows the processes of (1) Generation (Οφ + Gd) (2) recombina­
tion (Un) and (3) charge extraction.
Impedance spectroscopy 465

If we restrict our attention to a steady-state condition, Eq. (12.11) reduces to:


(12.12)

Comparing Eqs. (12.1) and (12.12), the photocurrent generated during short-
circuit becomes:
(12.13)

The total generation per unit area, LG<s> is proportional to the incident ligh
intensity, Εϋφ = ηορίΦ0, where T7opt is an optical quantum yield that depends on the
properties of absorption of the radiation by the solar cell. We also obtain that:
(12.14)

Consequently, in the ideal solar cell model with unit collection efficiency the
dark reverse current corresponds to the extraction of the carriers generated by the
thermal surrounding radiation.
We already appreciate that the ideal diode model of a solar cell states that a
constant current is drawn out of the cell, namely j s c +jd, which corresponds to all the
electron carriers generated in the semiconductor. In addition, the recombination term
produces a current in the opposite direction. At high forward bias the recombination
term dominates and bends the j-V curve, as indicated in Figure 12.3(a). Note that this
ideal model does not contain any trace of diffusion whatsoever. The only element
required in order to obtain the diode model is to state that the contacts are selective,
and extract only one carrier at each side, as indicated in Figure 12.4.
Another step for converting the conservation equation into a j-V characteristic is
to relate the carrier density, n, to the applied voltage, V, by introducing Fermi levels.
We assume the extended states for electrons at the level Ec (conduction band edge),
with an effective density, Nc. With respect to the electron Fermi level EFn, we have:
(12.15)
and considering the dark (equilibrium) Fermi level E¥0,
(12.16)

we obtain
(12.17)

The voltage, V, is measured at the selective contacts, and corresponds to the dif­
ference in Fermi levels of carriers at the contacts. If the contacts are ideally reversible
[12.33], each contact separately equilibrates with the Fermi level of electrons, EFn,
and holes, EF . This gives:
(12.18)
For a p-semiconductor, the holes in the Fermi level remain at the dark equilib­
rium level, EFp = £ F0 , and Eq. (12.18) can thus be written:
(12.19)
466 Dye-Sensitized Solar Cells

In consequence:

(12.20)

Using the linear recombination of Eq. (12.5) in Eq. (12.12), and applying the
Boltzmann statistics indicated in Eq. (12.20), we obtain the diode equation (12.1) with
m = 1. However, if we assume a nonlinear recombination model, more general than
the one used previously
(12.21)

we obtain the general diode equation with m = 1/ß. Here, Eq. (12.21) is written as a
purely empirical law, but its origin is further discussed below.
It should be noted that the recombination mechanism has a major impact on the
shape of the j-V curve, especially on the FF. Therefore also on the solar cell conver­
sion efficiency. In fact, as we have shown with the above model, for ideal selective
contacts, the diode ideality factor m is entirely determined by the bulk recombination
mechanism. This point has been discussed in Ref. [1].

12.3 INTRODUCTION TO IS METHODS

In general, IS is applied to a system with electrical contacts. It consists of a measure­


ment of the ac electrical current, 7(ω), at a certain angular frequency, ω, when a cer­
tain ac voltage, ν(ω), is applied to the system, or vice versa, a measurement of V(co)
at an applied Ι(ω). The impedance is:

(12.22)

The symbol x over a quantity x indicates that x is:

(1) the complex amplitude of a sinusoidal (ac) perturbation of x and

(2) a small perturbation.

The "smallness" of x is required in order to obtain the linear impedance in Eq.


(12.22), i.e., Ι(ω) is linear with respect to V(oo\ or vice versa, so that Ζ(ω) is inde­
pendent of the amplitude of the perturbation. In modeling work, this is ensured if the
absolute value of x is much lower than that of the steady state quantities x, y, ... In
practice this means that the amplitude of the voltage must be on the order of several
mV. However, in certain situations, e.g., close to a phase transition, a small perturbation
of the voltage induces very large variations of the charge or current, and the conditions
of linearity must thus be carefully inquired.
During an impedance measurement, the system is (ideally) kept at a fixed steady
state by imposing stationary constraints such as the dc current, illumination intensity,
etc., and the Ζ(ω) is measured by scanning the frequency at a multitude of values
Impedance spectroscopy 467

/ = ωΙΙττ, typically over several decades, i.e., from mHz to 10 MHz, with 5-10 meas­
urements per decade. At each frequency the impedance meter must verify that the Ζ(ω)
is stable. At low frequencies, this takes a considerable amount of time, i.e., stabilizing
a measurement a t / = 10 mHz consumes minutes. Nevertheless, measurements at low
frequencies are often important in order to make sure that one is approaching the dc
regime, as further explained below. A judicious selection of the frequency window of
measurement is therefore necessary, and this is often aided by experience.
In addition to scanning the frequencies, it is usually very important to deter­
mine the IS parameters at various conditions of steady state. This is the key approach
in order to relate the measurement to a given physical model. At each steady state the
Ζ(ω) data is related to a model in the frequency domain, which is usually represented
as an equivalent circuit. By modifying the steady state, the change in impedance
parameters (resistances, capacitances, etc.) can be monitored in relation to the physi­
cal properties of the system. Since the impedance measurement takes a considerable
amount of time, the steady state often changes along the impedance measurement,
and precautions should be taken to avoid a serious drift of the parameters. In par­
ticular, care should be taken with unintentional changes of temperature in solar cells,
since this introduces additional and unwanted variations of the parameters.
Note that, at each steady state, a full scan of frequencies is necessary. Thus
many steady state points imply a long measurement, perhaps over an entire day.
However, data that do not cover different steady states may in some cases be of little
value, particularly if there is uncertainty regarding the meaning of the parameters,
and especially when comparing different solar cell devices. It is also important to
verify the true significance of parameters by material variations of the samples, e.g.,
to confirm the correlation of a transport resistance with the reciprocal length of the
sample. The extent to which these approaches must be judiciously realized depends
on the preliminary knowledge and experience of the particular system.

12.3.1 Steady state and small perturbation quantities


As an example of the relationship between the ac impedance and steady-state quan­
tities, we discuss a characteristic experiment on a solar cell using the ideal model
outlined in Figures 12.3 and 12.4. We choose a certain point of bias voltage, V0, with
the associated current density, j0. At this point, a small displacement of voltage V(0)
implies a change of current j(0). The value ω = 0 in parenthesis indicates that the
displacement is infinitely slow, i.e., V(0) and j(0) attain a value that is independent of
time. The displacement of the current and voltage is indicated in Figure 12.3(a) with
arrows.
For a solar cell with area A, the quotient of the small quantities gives:

(12.23)

In other words, the small quantities provide a derivative of the voltage with
respect to the current. This is the reciprocal of the slope of the j-V curve, which is in turn
the dc resistance of the solar cell Rdc (per area) under those particular conditions.
468 Dye-Sensitized Solar Cells

A similar process occurs if we measure the change in the electrical charge, <2,
under a perturbation of the voltage. The quotient is a capacitance:

(12.24)

In general, the parameters obtained by IS are related to derivatives of the steady


state variables describing the system, i.e., IS provides the differential resistance, dif­
ferential capacitance, etc. However, we usually omit the specification of "differential"
in the context of IS as it is implicitly assumed.
It is useful to observe that, since Rdc is the reciprocal of the slope of the cur­
rent density-potential curve, Figure 12.3(a), knowledge of Rdc at several points allows
us to construct the full curve, provided that a single point of the curve is known (for
example, the value of j sc ):

(12.25)

Therefore, understanding the different elements that determine Rdc is a key step
in order to analyze the factors governing the efficiency of the solar cell.
From the steady state characteristic, we can only derive Z(0), i.e., the imped­
ance at the frequency ω = 0. However, in order to understand the operation of the solar
cell we wish to know the origin of Rdc in terms of the internal processes occurring in
the device: transport of charges, accumulation at certain points, recombination of car­
riers, and so on. Eventually, we are interested also in the dynamic behavior of the solar
cell, i.e., how it responds with time to a certain perturbation.
One way to obtain the dc parameters of the solar cell is to apply a certain model
of steady state operation. This can be done by an equivalent circuit that describes the
dc current distribution, including diode elements. This differs to ac-equivalent circuits
for IS spectra which are amply discussed below. In fact, since the diode is not a linear
impedance, it is not a differential element in the sense explained previously.
In particular a dc model, including an ideal diode, shunt resistance, rshunt, and
series resistance, rseries, is amply used in this context, see Figure 12.5 [12.70, 12.71].
This procedure normally assumes that rshunt and rseries are independent of the voltage

Fig. 12.5 The typical electrical model for inorganic semiconductor-based solar cells. The cur­
rent source accounts for the generation of electrons in the cell, the diode represents the recom­
bination characteristics, rshunt is a constant resistance accounting for charge losses crossing the
cell through the sides, and rseries also accounts for a constant resistance (contacts, wires, etc.)
Impedance spectroscopy 469

along the j-V characteristics. Such an assumption may work well in some classes of
solar cells such as monocrystalline Si solar cells. However, in other cases, especially
in devices including electrochemical processes such as in DSC, it is far from clear that
resistances remain constant, even at reverse voltage. Great care should be taken when
applying dc models to DSC, since one may impose a model that does not occur in the
device and the results of which may have little meaning.
We demonstrate later how to construct a dc model that is normally useful for the
analysis of DSC, but first we need to discuss the origin of the elements that appear in
equivalent circuits. To this end, we describe a much more powerful approach, apply­
ing IS, in order to obtain all the stationary and dynamic information concerning the
current-voltage behavior of the system.

12.3.2 The frequency domain


In general, the method of IS, consists in measuring the quotient in Eq. (12.22), for a
signal V(co) varying at different angular velocities. When the velocity, ω, is very slow,
we are close to steady state conditions and obtain exactly the dc resistance as indi­
cated in Eq. (12.23). However, when ω becomes faster, certain processes in the system
are unable to respond to the applied perturbation. Ζ(ω) therefore contains contribu­
tions from "things faster than" ω.
By scanning the frequency, we obtain a changing response (the impedance spec­
trum) that can be treated by several methods (analytical, numerical, and most impor­
tantly, by visual inspection of its shape) in order to provide a detailed physical picture
of the dynamic properties of the system. In particular, it is essential for solar cell
applications that this method renders it possible to dissect the steady-state response
into its elementary components. A vivid explanation of the physics and an interpreta­
tion of the electrical magnitudes in the frequency domain for dielectric materials is
given in the book Dielectric Relaxation in Solids by A.K. Jonscher [12.72].
One may wonder why one should use so many different angular frequencies in
the measurement, when the same processes can be probed by time transients, i.e., by
applying a voltage step and monitoring the subsequent evolution towards equilibrium.
This way, the fast and progressively slower processes in the system can be observed,
in a similar fashion as by the variation of the frequency of the perturbation.
Indeed, time transient methods are very important experimental tools, and
mathematically, small-amplitude time transients contain the same information as the
small-frequency linear impedance. Both are related by a Laplace transform. Indeed,
when the decay of the system is governed by a single process (usually an exponential
decay, with a characteristic time constant r), IS and time transients are equally valid
approaches. The difference arises when the response is composed of a combination of
processes. It then turns out that it is much easier to deconvolute the response, in terms
of models, from the spectroscopic response Ζ(ω) as opposed to from the featureless
time-dependent signal.
As another example of the advantages of the frequency domain, let us consider
the kinetic response of the capacitance that was derived for equilibrium conditions in
Eq. (12.24). In the time domain, we apply a small step of the voltage V(t) = AV · u(t),
where u(t) is the unit step function at t = 0, and we observe the consequent evolution
470 Dye-Sensitized Solar Cells

of the charge Q(t) that passes to the system. However, it is generally not feasible to
measure a charge transient, and we thus need to observe the current transient, I(t), and
perform an integration:

(12.26)

When considering this process in the frequency domain, we use the variable
s = ίω, where / = . The Laplace-transformation of a function f(t) to the frequency
domain is defined as:

(12.27)

and the application of the transform to Eq. (12.26) gives:

(12.28)

We now introduce a frequency-dependent capacitance that generalizes Eq. (12.24)

(12.29)

Here, 0*(ω) is a function of the frequency, and it coincides with the static dif­
ferential capacitance C at ω = 0. By applying Eq. (12.22), we obtain from (12.28):

(12.30)

This result demonstrates the straightforwardness in resolving the small step-


charging experiment provided that the impedance is known. We observe in Eq. (12.30)
that the conversion of impedance data to capacitance turns out to be a very simple
operation. This simplicity of conversion between very different electrical magnitudes
appears as a result of the convenient properties of complex numbers, as well as due
to the fact that, in the frequency domain, derivatives and integrals are constituted of
arithmetic operations involving s.
Switching the data between representations is a very useful tool of analysis in
IS. The most frequently used functions are described in Table 12.6 [12.73] indicating
also the separation of the magnitudes in their real and imaginary parts.

12.3.3 Simple equivalent circuits


Many measurements of IS in electrochemistry and materials devices can be described
by equivalent circuits composed of combinations of a few elements that are indicated
in Table 12.7. Equivalent circuits are formed by connecting these and other elements
by wires, representing low resistance paths in the system. Two elements are in series
when the current through them is the same, whereas they are in parallel when the
Impedance spectroscopy 471

Table 12.6 Impedance representations.

Denomination Definition* Real and imaginary parts

Impedance Ζ(ω) Ζ=Ζ' + ίΖ"

Admittance Υ(ω) = ~^— Y=Y' + iY"


Ζ(ω)
Z
* Ä=
tano "

Phase angle
Z'
Complex capacitance C*(<w) = — l - — C* = C' + iC"
ίωΖ(ω)
Conductivity σ*(ω) = er* = σ' + ίσ', cr'(O) = σ
ΑΖ{ω)
Complex dielectric constant ε*(ω) = LC*(œ)/A ε* = ε + ιε
er* = ζ'ωε*

Complex electric modulus Μ*(ω) = — ^ M* = M' + /M"


ε(ω)
*L is the length of the sample, A is the area.

Table 12.7 Basic ac electrical elements.

Denomination Symbol Scheme Impedance

Resistance
Capacitance

Inductor

Constant phase element (CPE)


Qn

voltage acting on them is identical. Using Kirchhoff rules, we add the impedances
for two elements in series and the resulting impedance is an equivalent description of
the initial connection (under an applied voltage, it produces the same current as the
combination that it replaces). For elements in parallel, we add the admittances (or the
complex capacitances) to form the equivalent impedance.
A first example of an equivalent circuit is the R\C\ series combination. From
the impedance

(12.31)

we obtain the complex capacitance

(12.32)
472 Dye-Sensitized Solar Cells

Here, the relaxation time is defined as


(12.33)
Let us look more closely at the meaning of the relaxation time, r b in relation to
the response of the system in the time domain. We consider the type of measurement
commented before, in which a change of voltage, AV, is applied at time t = 0, and for
which the subsequent evolution of the electrical current is monitored. In the frequency
domain, the step voltage V(t) = AV · u(t) has the expression:

(12.34)

and the electrical current can be written:

(12.35)

By inverting Eq. (12.35) to the time domain, we obtain:

(10.36)

In general, the process described by Eqs. (12.32) or (12.36) is an elementary


relaxation with the characteristic frequency:

(12.37)

The plot of the complex capacitance is shown in Figure 12.8(a). The capaci­
tance displays an arc from the dc value C*(0) = Cx to the high frequency value.
The top of the arc occurs at the characteristic frequency of the relaxation ωχ. The
impedance, shown in the complex plane in Figure 12.8(b), forms a vertical line. This
is a "blocking" circuit, since the impedance of a capacitor is oo at low frequency,
which effectively constitutes an open circuit connection, thus preventing the dc cur­
rent from flowing. However, the impedance of the capacitor decreases as the fre­
quency increases, and at very large frequencies, with respect to ωί9 the capacitor
indeed becomes a short-circuit. Consequently, there remains only the resistance Rh
The impedance of a resistor is the same at all frequencies, hence the vertical line in
Figure 12.8(b). The arc in Figure 12.8(a) is a manifestation of an elementary relaxa­
tion process that corresponds to an exponential decay in the time domain, indicated
in Eq. (12.36).
Another important example of an equivalent circuit is the RC parallel combina­
tion, depicted in Figure 12.9. The admittance of the combination is here:

(12.38)
Impedance spectroscopy 473

Fig. 12.8 Representations of the impedance of an equivalent circuit. Ri = 1 kQ, Cx = 1 mF,


Ti = 1 s. The thick arrows indicate the direction of an increasing angular frequency, ω.

With the addition of a series resistance, R2, we obtain the circuit shown in
Figure 12.9. The impedance is:

(12.39)

The complex impedance plot is shown in Figure 12.9(a). The parallel RC forms
an arc in the complex plane which is shifted positively along the real axis by the
series resistance, R2. As we remarked before, the capacitor can at zero frequency be
substituted by an open-circuit connection. In contrast to Figure 12.8, we observe in
Figure 12.9 that this is a circuit with dc conduction determined by the low frequency
intercept, Z(0) = Rdc = RX+ R2.
In Figure 12.9(a) the three plots correspond to a variation of the parallel resist­
ance, which implies a change in the characteristic time, τχ =RXCX. In the complex
plane, we readily infer the structure of the circuit from the shape of the spectra, but fre­
quency values and time scales cannot be directly read. To this end, it is useful to apply
the plot with respect to frequency (sometimes termed a Bode plot). Figure 12.9(b)
shows the transition of the resistance from the low frequency (Rdc) to the high fre­
quency value (R2). This high frequency value occurs due to the fact that the capacitor
impedance disappears at very high frequency, Ζ(ω = ©o) = 0, thus shunting the parallel
resistance. Another representation often used to display the characteristic frequencies
474 Dye-Sensitized Solar Cells

Fig. 12.9 Representations of the impedance of an equivalent circuit. R\ takes on values 5, 4,


2 kQ, Ci = 10 mF, n = 50, 40, 20 s, Ä2 = 1 kQ.

is the phase angle. Figure 12.9(c) shows that the peak of the phase angle moves to
higher frequencies when τγ decreases.
In measurement of material systems, it is rather frequent for the IS response
to be composed of the combination of several processes. The time constants, and the
connection of the elements describing such processes, depend on the internal structure
of the system. A primary aim of the data analysis is to identify the contribution of
separate relaxation processes in the frequency response of the system and such aim is
greatly assisted by picking the appropriate form of data display. In IS measurement
Impedance spectroscopy 475

we obtain the data, and such data can be transformed as desired between the various
representations of Table 12.6.
As mentioned in a previous section, the most critical information concerning
solar cell device operation in stationary conditions relates to the separation of resist­
ances. However, in IS, capacitances also play a crucial role, since different elements
with similar resistance provide very distinct spectral features if their associated capac­
itances differ sufficiently in magnitude. The capacitance is, therefore, a key to the
understanding of the origin of the measured resistances.
Figure 12.10 shows the example of a system composed of several relaxations rep­
resented by two series of RC circuits connected in parallel. This circuit is relevant for the
analysis of multiple-trap systems in electronic materials [12.74, 12.75]. The inspection
of the complex impedance plane in Figure 12.10(b) only shows the blocking response
at low frequencies and an additional feature at high frequency. For a blocking circuit, it
is natural to analyze the capacitance, and the plot of the capacitance components with

Fig. 12.10 Representations of the impedance of an equivalent circuit. R\ = 1 kQ, C\ = 5 mF,


Tl = 5 s, R2 = 0.1 kQ, C2 = 1 mF, r 2 = 0.1 s.
476 Dye-Sensitized Solar Cells

respect to the frequency, Figure 12.10(c), usually reveals a great deal of information.
In Figure 12.10(c), we observe two plateaus of the real part of the capacitance which
clearly indicate two distinct relaxation processes. These relaxations are manifested in
the peaks of the loss component of the capacitance, C. When increasing the frequency,
each peak of C indicates the occurrence of a relaxation and a consequent decrease of
the capacitance [12.72]. Such features can also be observed in the complex capacitance
plot in Figure 12.10(a), demonstrating separate arcs for the two relaxations.
Let us consider in more detail how to obtain the parameters of a given IS data
set. The main method consists in fitting by least squares methods using an equivalent
circuit software that is available in many kinds of measuring equipments. However, the
fitting process requires the assumption of a given equivalent circuit, and sometimes, in
addition, the input of reasonable trial parameters. As we have mentioned before, the
inspection of the data set in several complementary representations usually provides a
good hint of the equivalent circuit structure, at least in the less complex cases. Another
useful approach is to read the parameter values directly from the data representation,
e.g., resistances and capacitances of separate contributions. How to perform this has
already been discussed in the examples of Figures 12.9(b) and 12.10(c). However, the
values of capacitance or impedance in a certain frequency domain can be influenced
by the whole equivalent circuit. So, to obtain the circuit parameters, there is often
no substitute for integral data fitting. Separately treating part of the spectral data is a
valuable resource, but one that should be used with care.
For instance, in Figure 12.9(a) we observe that the impedance displays a verti­
cal line when approaching the dc limit. Therefore, at low frequency, Figure 12.9(a)
can be simply described by RC parallel combination. The low frequency resistance is
clearly given by Rdc. But what should be used as the low frequency capacitance C\p. It
cannot be Cu otherwise the arc would finish at the origin of Figure 12.9(a), which it
does not. In general, it is very useful to obtain the impedance formula in a restricted
frequency domain, and the method is demonstrated with this example.
First, from the expression of the impedance in Eq. (12.39), we find the low
frequency limit, which gives:
(12.40)

This last equation does not correspond to any recognizable combination of cir­
cuit elements. In fact, we seek a parallel combination, which should provide a good
description of the data in Figure 12.9(a) at low frequencies. Consequently, we trans­
form Eq. (12.40) to the admittance, maintaining the first order approximation in ω,
with the result:

(12.41)

In Eq. (12.41), we readily recognize the parallel RC admittance formula. The


low frequency capacitance is:

(12.42)
Impedance spectroscopy 477

The capacitance therefore depends on the resistances of the original circuit. This
result is quite natural, since the capacitance relates to the reciprocal of the impedance
(see Table 12.6), and the latter is greatly influenced by the series resistance. However,
the result in Eq. (12.42) cannot be inferred without a proper calculation.
Let us continue with the analysis of the effect of different types of equiva­
lent circuit elements. While the combination of resistances and capacitors provides a
spectrum that remains in the first quadrant of the complex impedance plane, it is not
uncommon to find that the data cross to the fourth quadrant. One reason for this is
the inductance of the leads, which very frequently causes a tail at high frequencies in
which the spectrum crosses the real axis. A different feature is often found in several
types of solar cells at low frequency; consisting in a loop that forms an arc in the fourth
quadrant [12.76]. One of the representations of this effect is a series RL branch com­
plementing the RC circuit of Figure 12.9. The model is shown in Figure 12.11, and
the total admittance has the value

(12.43)

The low frequency limit of Eq. (12.43) is written

(12.44)

Eq. (12.44) shows that, when R3 is small, the capacitance becomes negative at
low frequencies, i.e., C = -C N with the value

(12.45)

The spectra with both a positive and negative low frequency capacitance are
shown in Figure 12.11(a). If R3 < (L3/Ci)1/2, the impedance traces a low frequency arc
in the fourth quadrant, otherwise, the impedance remains in the first quadrant. The
intercept of Z with the real axis (i.e., the transition of C\œ>) to negative values) occurs
at the frequency

(12.46)

In the capacitance vs. frequency representation, Figure 12.11(b), the presence of


the inductor appears as the negative contribution that becomes more negative towards
lower frequencies. At high frequencies, the plot is dominated by Ch whereas at lower
frequencies, the circuit capacitance starts to decrease due to the inductive effect. At ω ΝΟ
it shows a dip at the transition from positive to negative values, after which the absolute
value increases towards lower frequencies, until it saturates at the value - CN.
As a final example of the simple equivalent circuits, we consider the presence of
a Constant Phase Element (CPE) as shown in Figure 12.12. The normal application of
478 Dye-Sensitized Solar Cells

Fig. 12.11 Representations of the impedance of an equivalent circuit. Ri = 1 kQ, C\ = 1 mF,


L3 = 1 kH, R3 = Rxla, a varies as indicated.

a CPE is to describe a capacitive process that presents a certain frequency dispersion.


The latter occurs when the CPE index n departs from 1. In fact, the pure capacitance
response with n = 1 is very rare, and it is often necessary to use CPEs with n < 1 in
fitting of data [12.46]. Despite such a widespread occurrence, a general origin for CPE
responses in terms of a unique physical process has not been identified. CPE is related
to systems that show some kind of self-scaling, either of geometric origin (such as
fractal electrodes [12.77]) or dynamical origin (like in certain multiple trapping sys­
tems [12.75]). Due to self-scaling properties of the CPE response, it is normally diffi­
cult to identify the specific factor causing the dispersion, and CPE should be regarded
as a useful and often indispensable tool for data description.
When index n decreases, the modification of the capacitive response becomes
rather large, whereas the RQ arc becomes progressively depressed, as shown in
Impedance spectroscopy 479

Fig. 12.12 Representations of the impedance of an equivalent circuit. Rx = 2 kQ, Cx = 10 mF.


sn_1, n varies as indicated, and R2 = 0.1 kQ.

Figure 12.12(a). The CPE also gives rise to a significant deceleration of the response.
Figure 12.12(b) shows that the transition from a low to high frequency resistance
of the capacitor response is completed in less than two decades of frequency, while
for n = 0.6, it requires more than four decades. Consequently, the characteristic
frequency presents an important reduction as n decreases, according to the expres­
sion [12.46]

(12.47)

The foregoing discussion has shown that equivalent circuit representations are
a very powerful resource for the inverse problem that is usually a main task in IS data
treatment: to establish an impedance model from a set of data. Importantly, equivalent
480 Dye-Sensitized Solar Cells

circuits render it possible to visualize the structure of the model and to separately treat
data portions in certain relevant frequency windows. However, equivalent circuits are
by no means necessary in order to establish a physical model; what is needed is an
impedance function, in any of all its possible analytical representations.
It should also be mentioned, that not all complex functions of frequency are
valid impedance responses. The complex function Ζ(ω) must obey causality condi­
tions (i.e., the stimulus must precede the response), which imposes analytical con­
straints known as Kramers-Kronig transforms [12.72]. These transforms enable the
construction of the real part of Ζ(ω) provided that the imaginary part is known at
all frequencies, and vice versa. Using equivalent circuit elements, such as those of
Table 12.6, ensures that the resulting model obeys the Kramers-Kronig relations.

12.4 BASIC PHYSICAL MODEL AND PARAMETERS OF IS IN


SOLAR CELLS

12.4.1 Simplest impedance model of a solar cell


In the process of obtaining physical information from IS data, it is necessary to relate
the observable equivalent circuit elements with the system properties. As mentioned
before, equivalent circuits are a useful tool for interpretation, and the significance
attached to the circuit elements, the potential in the circuit, etc., may be quite different
from the standard physics textbook examples.
This is particularly the case in the analysis of solar cells. Note that the ac-
equivalent circuits that we have discussed are composed of passive elements (i.e.,
resistances and capacitances). It is common to interpret the flow of charges in cir­
cuits in terms of the mechanistic view of the drift of charges in an electrical field
caused by potential differences. This image is also very popular for explaining the
photovoltaic action, e.g., in a p-n junction, in terms of an electric field that sends
oppositely charged carriers in different directions. However, a solar cell is a kind
of battery, i.e., an element producing an electromotive force, and such an element
cannot work with electrostatic voltage differences alone. According to Volta's idea,
the electromotive force is an nonelectrostatic action on charges in conductors that
causes unequal charges to separate and remain separated [12.78]. We thus wish to
obtain the internal ac-equivalent circuit of a solar cell using only linear elements
associated to a small signal ac perturbation, with emphasis on the interpretation
of the elements that make it work as a device for the production of electricity. The
key approach for useful reading of ac-equivalent circuits of DSC, is that potentials
in the circuit represent an electrochemical potential of electrons (or holes) in the
actual device.
To clarify this, we start with the simplest model of a solar cell, discussed above
in Section 12.2.2, which contains the necessary elements without complications
of carrier transport, specific features of selective contacts, etc. We calculate the IS
response of the solar cell of Figure 12.4 [12.35], corresponding to the application of a
small ac electrical perturbation.
Impedance spectroscopy 481

It was demonstrated before that the dynamic response of the simple solar cell
model in Figure 12.4 was determined by the equation:

(12.48)

where G = G<p + Gd is the carrier generation rate. In order to calculate the IS response,
we need to combine two approaches: (1) All physical quantities are composed of a
stationary part (e.g., n) and a small perturbation part that varies with time. (2) We
must reduce all the dependencies implicit in Eq. (12.48), explicitly to voltage, so that
the result becomes an impedance.
For example, the carrier density dependence on time takes the form:
(12.49)

The variation of voltage applied in the solar cell produces a variation of the
electron Fermi level, which changes according to:
(12.50)

where φη is the small perturbation voltage. However, by Eq. (12.15) there is a unique
dependence of n on EFn. Thus:
(12.51)

Expanding Eq. (12.51) to first order, we obtain:

(12.52)

The derivative in Eq. (12.52) appears recurrently in solar cell theory and
requires a special denomination. We introduce the chemical capacitance, a thermody-
namic quantity that reflects the capability of a system to accept or release additional
carriers with density N{ due to a change in their chemical potential, μ{ [12.35, 12.79].
In general, for a volume element that stores chemical energy due to a thermodynamic
displacement, the chemical capacitance per unit volume is defined as:

(12.53)

More generally, Eq. (12.53) employs the electrochemical potential, that coin­
cides with the electron Fermi level [12.60]. Thus, the chemical capacitance for con­
duction band electrons is [12.351:

(12.54)

The macroscopic capacitance for a film of thickness L, area A and porosity p is


written as:
(12.55)
482 Dye-Sensitized Solar Cells

where the quantity LA(1 -p) is the volume of the film Vf. Using Eqs. (12.49), (12.52)
and (12.54), we arrive at the relationship between the small perturbation of carrier
density and voltage:

(12.56)

Next, we expand the recombination term in Eq. (12.48) and obtain:

(12.57)

Finally, the current can be expressed as

(12.58)

When we insert the different expanded expressions into Eq. (12.48), we first
obtain a time-independent equation that has already been discussed, (12.12), and
which gives the stationary condition of the solar cell according to bias voltage and
illumination. In addition, the time dependent terms provide a new equation that takes
the form:

(12.59)

where the recombination resistance per unit volume is given by:

(12.60)

Note that Eq. (12.60) can be represented by the following expression:

(10.61)

and the macroscopic recombination resistance, which corresponds to the reciprocal


derivative of the recombination current with respect to voltage, is written as:

(12.62)

The fundamental parameter describing recombination, however, is the recom­


bination per unit of effective internal area (Aeff) r[ = AQiiRr =LA(l-p)hRr =hrr,
where h is the ratio between the effective area and the volume of the Ti0 2 film, i.e.,
h=AeiiILA{\-p).
We remark that the carrier generation terms are absent from Eq. (12.59), since it
is only possible to modulate electrical injection of carriers in IS; the situation is differ­
ent in light-modulated techniques, as explained by Peter and Hagfeldt in Chapter 11.
Impedance spectroscopy 483

The structure of the impedance model can be inferred directly from Eq. (12.59)
[12.35]. φη can be viewed as the potential in an equivalent circuit (but one should
remember that it is physically the electrochemical potential!). Then, Eq. (12.59) is
Kirchoff 's rule for current conservation. The first term is a capacitive current, the sec­
ond is an ohmic current through the resistor Rn and the third is the extraction current.
Note that the two first currents do not represent transport currents (i.e., an ensemble
of carriers moving in a certain direction in space), but rather the rates of creation and
destruction of conduction band electrons.
In order to calculate the impedance, we apply the Laplace transform (d/dt —» ίω)
in Eq. (12.59) and use the definition in Eq. (12.22):

(12.63)

Eq. (12.63) clearly corresponds to the parallel combination of the chemical


capacitance and recombination resistance, which is the minimal IS model of a solar
cell. Figure 12.13(b) shows the equivalent circuit corresponding to the basic solar cell
scheme of Figure 12.13(a).
In Figure 12.13(b), we can observe that the chemical capacitance is a neces­
sary element of the solar cell: it produces a voltage (associated to a splitting of Fermi
levels) by the creation of excess carriers from photons. An important message in
Figure 12.13 is that the recombination resistance needs to be be large, as this will
allow carriers accumulated in the capacitive element to flow through the external
circuit when returning to the equilibrium situation. We point out that the recom­
bination resistance in Figure 12.13(b) corresponds to the diode in the dc circuit of
Figure 12.5.
Figure 12.13(b) also displays the special structure of connection of the R and
C elements by selective contacts which is implicit in the derivation of the result in
Eq. (12.63). This connection is essential in order to channel the carriers in the desired
direction. An example of the failure of selective contacts is shown in Figure 12.13(c).
Electrons and holes meet directly at the left contact, producing an internal short cir­
cuit. Such a device cannot produce a photovoltage.
It should also be recognized that, in contrast to electrochemical batteries and
capacitors, there is in solar cells always an electrical connection between the outer
electrodes via the internal resistance, rr. In fact, the solar cell works by promotion of
carriers from a low to a high energy level, with the energy of the photons [12.33], and
such energy levels are separately connected to the outer electrodes. Since the excita­
tion is possible, the converse process, which is the decay from a high to a low energy
level by radiative recombination, must also be possible. This is the most favorable
case of the recombination resistance, which is unavoidable, as it is an intrinsic com­
ponent of the photophysical process causing the solar cell to produce useful work. In
this sense, we regard Figure 12.13(b) as the minimal model.
Nevertheless, while certain recombination processes are unavoidable in the
solar cell, additional sources of recombination are detrimental to the performance. For
example, in Figure 12.13(c), a strong recombination at the left contact produces a low
484 Dye-Sensitized Solar Cells

Fig. 12.13 Scheme of injection of electrons (1) and holes (3) as well as recombination (2)
processes of a solar cell, induced by the application of a potential.

internal resistance, and this in a process that does not contribute at all to the carrier
generation. Such a behavior must be regarded as a failure of the device. In fact, reduc­
ing surface recombination is the most critical step in the preparation of high efficiency
industrial silicon solar cells [12.80]. In general, equivalent circuits illustrate a main
point with regard to solar cell operation: the dc current predominantly follows the path
of least resistance. Therefore, low resistances in parallel to the chemical capacitance
reduce the output power.
The ideal model provides a very useful reference for understanding IS results of
solar cells. However, it should be emphasized that one of the goals of hybrid nanos-
tructured organic-inorganic solar cells is to obtain low-cost photovoltaic devices. For
this reason, there are additional elements contributing to the photovoltaic conversion
process. A key feature rendering IS attractive is that it can be applied in full devices
Impedance spectroscopy 485

and indicate the main limitations to photoelectrical performance. We will in subse­


quent sections progress towards a full realistic model for DSC devices, but a simple
example may be illustrative.
The common way to construct a selective contact for holes is by using a hole-
transport material that readily conducts hole carriers and blocks electrons. This is shown
in Figure 12.14. However, the organic conductors, which are able to penetrate the pores
of T i 0 2 nanostructures, generally posses a limited carrier mobility. Since the conduc­
tivity is low, a gradient of the hole Fermi level is required to inject, as in Figure 12.14
(a), or to extract the holes across the layer. Comparing this case with the ideal selective
contact in Figure 12.13(a), the difference is that, in the latter case, the extraction of holes
represents no cost at all in terms of the Fermi level gradient. We have mentioned above
that parallel resistances should be large, and, from the present example, we appreciate
that series resistances must be relatively small to avoid power losses.
The Fermi level drop for hole transport in Fig. 12.14(a) implies a "potential
drop" in the equivalent circuit of Fig. 12.14(b), with an associated impedance which
is related to hole diffusion. Therefore, problems in the performance of contacts, or
transport layers, can be detected with IS measurements. To do so, we must be able
to separate, in the IS data, according to the previous example, the contribution of
the recombination resistance and the transport layer resistance. This will largely
depend on the values of capacitances of the two elements. As mentioned before, the

Fig. 12.14 (a) Energetic diagram of a solar cell and scheme of the processes of injection of
electrons (1) and holes (4), recombination (2) and transport losses in the hole-conducting layer
(3). (b) Equivalent circuit for the ac electrical perturbation.
486 Dye-Sensitized Solar Cells

interpretation of the capacitance is a major tool for identifying the physical origin of
processes observed in IS measurements.

12.4.2 Measurements of electron lifetimes


It is interesting to explain in more detail the relationship between the equivalent
circuit elements describing the solar cell IS response and the electron lifetime. In
order to describe the IS behavior, we have considered in Eq. (12.63) an experiment
relating voltage to an electrical current measurement. However, we can employ the
general dynamic equation (12.59) in experiments in which we apply a perturbation
and let the system decay by itself [12.54, 12.57]. Since no current is extracted, we
obtain:

(12.64)

Eq. (12.64) describes, for instance, the exponential decay of a small step of
excess carrier concentration by recombination. From Eq. (12.64), the time constant of
the decay process, which we denote the response time, is:
(12.65)

In the model outlined above, this also gives:

(12.66)

With the normal assumption of a first-order reaction for direct electron transfer
from the conduction band, Eq. (12.5), we obtain simply r r = r 0 . In this simple model,
the lifetime is constant, and the response time and electron lifetime have the same
meaning. But in general, the lifetime can be dependant on steady-state conditions,
as is obvious in Eq. (12.66). In addition, in the presence of additional relaxation
processes such as trapping and release in localized electronic states, the response
time contains components due to kinetic delays in addition to the free carrier life­
time [12.40, 12.166].

12.5 BASIC PHYSICAL MODELS AND PARAMETERS OF IS


IN DYE-SENSITIZED SOLAR CELLS

12.5.1 Electronic processes in a DSC


A general view of the electronic and ionic processes occurring in a DSC is given in
Figure 12.15. With respect to the basic solar cell model in Figure 12.4, the sensitizer
in a DSC (molecular dye, inorganic quantum dot, etc.) is the absorber [12.33]. The
selective contacts to the absorber are formed, first, by an electron transport material
Next Page

Impedance spectroscopy 487

Fig. 12.15 A schematic energy drawing of the electron- and hole-transfer processes at the
metal-oxide (ETM)/dye layer/hole-transport material (HTM) in a DSC. The boxes indicate
available electronic states, EF0 is the dark Fermi level, and Ec is the lower edge of the metal
oxide conduction band. (1) Photoexcitation. (2) Electron injection from the dye LUMO to the
metal oxide. (3) Electron diffusion in ETM. (4) Electron injection to the TCO. (5) Hole-transfer
from the dye HOMO to the HTM HOMO. (6) Hole-diffusion in HTM. (7) Regeneration with
an electron from the counter electrode. (9) Electron transfer from metal oxide to HTM and to
(8) dye HOMO.

(ETM), which is the wide bandgap semiconductor nanostructure on top of a TCO,


where Ti0 2 is the archetypical semiconductor. The second selective contact is a
hole-transport material (HTM), which in original DSC is a redox carrier in a liquid
electrolyte. In fully solid devices, on the other hand, it is an organic hole-conductor,
as explained by Snaith in Chapter 6.
In IS, we do not directly monitor the photoinjection process, as explained above,
and the main attention is focused on electronic processes of electrons in the ETM (and
eventually holes in solid state HTM), which are described in Figure 12.16. The rea­
son why IS relates predominantly to electrons in the wide bandgap semiconductor,
is that the concentration of redox carrier in an electrolyte is very high (approaching
1020 cm -3 ). Moreover, it is hardly affected by the bias, whereas the electron concen­
tration changes by many orders of magnitude when the potential is displaced [12.81].
We can therefore monitor wide variations of the IS parameters related to electronic
processes, as discussed below.
Materials for hybrid solar cells based on low-cost semiconductors usually
include a large extent of electronic energy disorder, implying a wide distribution of
localized electronic states in the bandgap, as indicated in Figure 12.16 [12.61]. The
transport of electrons is usually described in terms of a classical multiple trapping
transport [12.61]. This model includes two classes of electronic states: the transport
CHAPTER 13

THEORETICAL AND MODEL SYSTEM


CALCULATIONS
Filippo De Angelis and Simona Fantacci

13.1 INTRODUCTION

Within today's global challenge to capture and utilize solar energy for a sustainable
development on a grand scale, dye-sensitized solar cells (DSCs) represent a particu­
larly promising approach to the direct conversion of light into electrical energy at low
cost and with high efficiency [13.1-13.6]. In these devices, a dye sensitizer absorbs
the solar radiation and transfers the photoexcited electron to a wide band-gap semi­
conductor electrode consisting of a mesoporous oxide layer based on nanometer-sized
particles. The concomitant hole is transferred to the redox electrolyte [13.7], typically
iodide/triiodide in solution or a hole-transporting material in the solid state.
Various ruthenium(II) complexes have been employed as dye sensitizers [13.8],
with carboxylic acid, dihydroxy, and phosphonic acid groups on the bipyridine lig-
ands anchoring the dyes on the surface of the oxide, typically Ti0 2 , nanoparticles.
The remarkable performance of the tetraprotonated [d^dithiocyanato)-Ru-fe(2,2'-
bipyridine-4,4/-dicarboxylate)] complex (N3) and its doubly protonated analogue
(N719), see Figure 13.1, has had a central role in significantly advancing the DSC
technology [13.9]. Fully organic sensitizers have also been developed due to their
higher molar extinction coefficient, as compared to Ru(II)-dyes, spectral tunability
and reduced environmental impact [13.10-13.13]. These dyes have, however, deliv­
ered consistently reduced performance as compared to their Ru(II)-N719 counterpart,
for which solar-to-electric power efficiencies exceeding 11 % have been reported
[13.14]. DSCs based on liquid electrolytes have received a substantial impulse to
industrialization, due to their high efficiency as well as low production and materi­
als costs. Despite this, higher conversion efficiencies need to be achieved in order to
obtain further progress. To this end, new sensitizers and a deeper understanding of the
interaction between the dye and the Ti0 2 nanoparticle are essential. Currently, there
is a tremendous impetus in the design of new and more efficient sensitizers. In certain
556 Dye-Sensitized Solar Cells

Fig. 13.1 Geometrical structures of [Ru(bpy)3]2+ (left) and of the N719 complex (right). TBA
counterions have been omitted from the N719 structure for the sake of clarity.

cases, molecular engineering together with advanced quantum chemical calculations


have been successfully employed [13.14-13.20].
The overall conversion efficiency (η) of a DSC is determined by the photocur-
rent density (/ph), the open circuit potential (V0CX the fill factor (ff) of the cell and the
intensity of the incident light (/s), according to:

(13.1)

It has been found that η depends markedly on several detailed features of the
sensitizer, such as the number of protons carried by the sensitizer's anchoring groups
[13.21, 13.22]. Indeed, upon adsorption, some of these protons may be transferred
to the Ti0 2 surface [13.23], and the resulting electric field assists electron injection,
favoring higher photocurrents. However, the proton-induced positive shift of the Fermi
level decreases the gap between the electrolyte redox couple and the Ti0 2 conduction
band, resulting in a lower open-circuit potential.
Theoretical calculations can be of great help in the design of new solar cell sen-
sitizers with improved characteristics [13.11-13.20], thereby providing a deep under­
standing of the character of the excited states involved in the absorption and injection
processes. A further step towards DSC optimization involves an effective modeling of
the electronic structure and optical properties of Ti0 2 surfaces or nanoparticles. It is
indeed important to evaluate the alignment of the dye-excited states with the semicon­
ductor conduction band. In doing so, one has to migrate from the molecular scale to
the nanoscale, and the adopted computational strategy should reflect this issue.
Clearly, the ultimate goal in the theoretical simulation of DSCs is the com­
putational modeling of combined dye/semiconductor systems, investigating the dye
adsorption mode onto Ti0 2 , the dye/Ti02 energy level alignment and electronic cou­
pling, the nature and localization of the excited states at the dye/semiconductor inter­
face, the effect of surface modifications and solvation on the electronic and optical
properties as well as the dynamics of electron injection. A further extension of these
studies should also include the electrolyte.
Theoretical and model system calculations 557

While theoretical and computational investigations of molecular dye sensitizers,


including those based on transition metals, are nowadays being successfully performed,
sometimes even by non-specialists in the field, the simulation of the dye/semiconductor
interface, on the other hand, is a much more complex problem, bridging computational
solid state physics/materials science and quantum chemistry. The latter thus requires a
multidisciplinary computational approach, and sometimes coding and implementation
of specific computational tasks. For this reason, a much more limited number of research
groups are performing such investigations and, to the best of our knowledge, nobody has
so far jointly addressed all the points raised above in a unique investigation.
The following is a review of the main theoretical, methodological and com­
putational results relevant to the simulation of DSCs. We begin with an introduction
on theoretical aspects followed by a review of the main results starting from the dye
molecules, both organic and inorganic, moving to Ti0 2 nanoparticles, and eventually
to the interacting dye/Ti02 systems.

13.2 THEORETICAL AND COMPUTATIONAL METHODS

The major difficulty in the theoretical and computational comprehension of nanos-


tructured materials resides in the complexity of the systems under investigation. The
complex interatomic interactions underlying nanoscale systems call for the use of
accurate computational techniques, while the large dimensions of these systems sub­
stantially limit the accuracy of the computational tools which can be employed. Even
computational tools demonstrating a reasonable compromise between accuracy and
computational overhead still have to tackle the inherent complexity of systems com­
posed of several hundreds (thousands) of atoms, which usually display a large number
of relevant geometrical configurations. The situation is even more severe if one con­
siders properties related to excited states. The accurate prediction by computational
methods of excited state properties of small to medium molecules still represents a
challenge in theoretical chemistry.

13.2.1 Density Functional Theory (DFT)


The development of ground-state theoretical tools rooted in Density Functional
Theory (DFT) has had a tremendous impact on theoretical and computational chem­
istry and materials science. The theory foundation dates back to 1964 and is repre­
sented by the Hohenberg-Kohn theorems which ensure that all the properties of a
physical system of interacting electrons can be described by means of the sole sin­
gle-particle electron density. The interactions among electrons are described in DFT
in the exchange-correlation (XC) functional, which includes all the exchange and
correlation interactions among the electrons. The resulting working equations, i.e.,
the Kohn-Sham (KS) equations, are mono-electronic and provide a set of molecular
orbitals which are used to construct the electron density. In practice, an approximate
XC functional is used, the quality of which, together with the basis set used to expand
the density, see below, fixes the accuracy of the calculation. The accuracy reached by
558 Dye-Sensitized Solar Cells

current XC functionals, especially hybrid ones, such as the renowned B3LYP [13.24],
including a certain amount of Hartree-Fock exchange, coupled to the reasonable scal­
ing of the computational cost, the increasing computer power of off-the-shelf PCs
and the availability of general-purpose quantum-chemistry (ADF [13.25], Gaussian03
[13.26], Dalton [13.27], and Turbomole [13.28]) or solid state physics (Quantum
Espresso [13.29], VASP [13.30], and Crystal [13.31]) has enabled also non-special­
ists of the field to routinely apply DFT to systems of medium dimensions.

13.2.2 Basis sets


As mentioned above, the electron density is built from the KS orbitals; these are in turn
expanded on a finite basis set for a practical solution of the KS equations. Two fami­
lies of basis sets exist: (i) localized basis functions, centered on the nuclei, and, (ii)
delocalized functions, typically plane waves (PWs), confined within a give volume.
The two families of basis sets are respectively related to computer codes stemming
from solid state physics, and quantum chemistry. Localized basis sets are the natural
choice for small medium molecular systems, for which the electron density is highly
inhomogeneous and mainly centered around the nuclei. Plane waves, on the other
hand, are perfectly suited to represent the electron density of solids and of periodic
systems, due to the inherent delocalized and periodic nature of these functions. PWs
codes are generally limited to non-hybrid exchange-correlation functionals. Hybrid
functionals are efficiently implemented in localized basis functions; however these
basis sets introduce a higher computational overhead.

13.2.3 The Car-Parrinello method


The possibility of performing DFT-based molecular dynamics simulations, by means of
the Car-Parrinello method [13.32], allows researchers not only to investigate the geometry
and electronic structure of complex systems but also to study the dynamics of chemical
reactions [13.33], to include the effect of thermal motion on the investigated proper­
ties [13.34, 13.35] and to optimize the geometry of extended systems characterized by
several local minima of the potential energy surface [13.36, 13.37]. The Car-Parrinello
(CP) method is a classical molecular dynamics scheme based on an interatomic potential
derived on the fly, i.e., for every nuclear configuration encountered during the system
time evolution, from DFT. In most standard implementations, the CP method employs a
PW basis set. Particularly efficient implementations of the CP method can be obtained
considering the so-called "ultrasoft" pseudo-potentials [13.38]; our benchmark calcula­
tions indicate that a saving of at least 2-3 times can be achieved at a comparable accuracy
[13.39]. Notably, CP-optimized geometries are usually as accurate as those obtained by
conventional quantum chemistry methods [13.33, 13.40].

13.2.4 Solvation effects


The inclusion of solvation effects is mandatory to provide a direct connection of the
calculated properties with the corresponding experimental quantities [13.15]. While
Theoretical and model system calculations 559

the explicit inclusion of solvent molecules is obviously the "exact" way of treat­
ing solvation effects, such a method involves a huge increase in the dimensions of
the system and the associated computational overhead. This approach has therefore
been limited to small/medium solutes [13.34, 13.41]. For this reason, the inclusion
of solvation effects is generally introduced by means of continuum solvation models
[13.42, 13.43] in which the solvent is treated as a structureless dielectric medium,
confined within a cavity defined by the molecular geometry. Typical implementations
of continuum solvation models available in commercial program packages include
the Polarizable Continuum Model (PCM) [13.42] and the COnductor-like Screening
MOdel (COSMO) [13.43].

13.2.5 Excited states


The Time-Dependent extension of DFT (TDDFT) [13.46] has become the method
of choice. This is due to its the accuracy coupled with its reasonable scaling with the
systems dimensions. TDDFT can be as accurate as correlated ab initio techniques
for the description of excited states [13.47], displaying a much lower computational
cost. TDDFT is still limited to excited states having a single-excitation character,
and various problems of current XC functionals have been highlighted for long-range
charge-transfer excitations [13.48] in which the starting and arriving orbitals of a
given transition do not overlap significantly. Nevertheless, TDDFT is currently suc­
cessfully applied to the study of organic molecules and systems containing transition
metal centers [13.49]. For these systems, upon inclusion of solvation effects and ther­
mal motion, TDDFT can yield excitation energies within 0.1-0.2 eV from experimen­
tal values [13.16, 13.17]. Moreover, considering the ground state geometry, TDDFT
can simulate absorption spectra. Non-equilibrium solvation models render it possible
to include the solvent response due to the excitation process [13.50]. Due to recent
implementations, TDDFT also enables the calculation of excited state geometries
[13.51]. This opens the way to calculations of emission spectra and excited state
dynamics. Furthermore, efficient procedures for calculations of dense spectra have
recently been resported [13.52].

13.2.6 Nonadiabatic method


The simulation of electron injection in DSCs can be performed in the framework
of Nonadiabatic Molecular Dynamics (NAMD). Degrees of freedom from both the
nuclear and the excited state are propagated in time from an initial condition, which
is usually taken to be the Frank-Condon point. Clearly, this procedure is perfectly
suited to simulate the electron injection process occurring in DSCs after light absorp­
tion. Atomic motions produce non-adiabatic electronic transitions between potential
energy surfaces corresponding to different excited states [13.53]. The whole NAMD
framework is extremely computationally demanding and various approximations have
been introduced to cope with the realistic systems. A typical approximation involves
using a semi-empirical Hamiltonian [13.54]. Alternatively, the KS orbital basis can
be employed, thus limiting the propagation to an unoccupied molecular orbital, rather
560 Dye-Sensitized Solar Cells

than to the coupled linear combination of occupied-to-virtual single excitations com­


posing the true excited state representation [13.53]. For extended systems, a usually
fair approximation concerns the use of a ground-state evolution of the nuclei [13.53,
13.54]. Obviously, all these approximations may lead to inaccurate time-scales of the
electron injection.
Most of the applications described in the following have been obtained with
reference to the theoretical/computational framework illustrated above. We therefore
limit the computational details to those strictly necessary for a thorough discussion of
the results of relevance for the DSC technology, refering the reader to the specialized
literature for further details.

13.3 DYE SENSITIZERS

This section reviews some of the most significant results obtained from the modeling
of organic and Ru(II)-dye sensitizers with direct application in DSCs. After a few
initial studies, there has during the last two-three years emerged a large amount of
literature on this subject. As a matter of fact, Ru(II)-polypyridyl complexes nowadays
represent a benchmark for current theoretical/computational methods.
Organic sensitizers have only very recently received attention for application
in DSCs. Due to the lack of systematic computational investigations on organic dyes,
numerous aspects concerning their photo-physical properties still need to be clarified.
Furthermore, some problems with "conventional" theoretical approaches have been
highlighted, and these will have to be properly addressed for a thorough understand­
ing of this extremely important class of dye-sensitizers.

13.3.1 Ruthenium(II)-polypyridyl sensitizers

Due to the success of the N3/N719 sensitizer in DSCs, most of theoretical/computa­


tional investigations have been initially devoted to this complex. Before we proceed,
however, we need to briefly mention the pioneering work by Daul, Baerends and
Vernooijs who, based on the X-ray geometry, performed a thorough DFT investigation
on the [Ru(bpy)3]2+ complex (Fig. 13.1) in vacuo back in 1994. They characterized the
electronic structure and the excited states of the complex, including spin-orbit-cou­
pling, in a very elegant and sophisticated way [13.55]. The lowest optical transitions
were correctly identified as being a of Metal-to-Ligand Charge Transfer (MLCT)
character, in agreement with the accepted experimental assignment [13.56].

13.3.2 Calculations on N3
The first theoretical investigation performed on N3 (and related complexes) dates
back to the work by Rensmo et al. in 1997 [13.57]. These authors investigated the
electronic structure of the N3 complex in the gas-phase considering a model geom­
etry, and analyzed the energy and character of the N3 molecular orbitals in relation
to the information on the density of states obtained by the experimental counterpart
Theoretical and model system calculations 561

of this study. Surprisingly, the calculated and experimental photoelectron spectrum


revealed that, conversely to [Ru(bpy)3]2+, the HOMOs of N3 were not pure Ru-t2g
states but rather resulted from the admixture of Ru-t2g states with sulfur lone pairs
and ΊΪ orbitals of the thiocyanate ligands. The LUMOs were, on the other hand,
expectedly assigned to π* orbitals of the functionalized bipyridine ligands, extend­
ing across the entire ligands, including the terminal carboxylic groups. Although it
was based on qualitative calculations, the observation of the presence of a thiocy­
anate character in the N3 HOMOs has important consequences for the functioning of
the complex in DSCs. Indeed, after injection of a photoexcited electron into the Ti0 2
conduction band, the dye is oxidized. Upon adsorption of the complex onto Ti0 2 , the
thiocyanate groups point towards the outermost region, which is that exposed to the
Γ/Ι3 electrolyte. As a result, a charge hole partially delocalized over the thiocyanate
ligands should promote the regeneration of the oxidized dye by the electrolyte redox
couple.
The first TDDFT study on the N3 complex, specifically on its tetradeproto-
nated form, was reported by Monat et al. in 2002 [13.58], while a semiempirical
ZINDO investigation on the analogous [Ru(4,4/-COOH-2,2/-bpy)2(Cl)2] complex
was reported in 2000 by Nazeeruddin et al. [13.59]. For the latter, the geometry was
optimized using DFT calculations. The TDDFT study by Monat et al. on N3 was
limited to the gas-phase and no geometrical optimization of the molecular structure
was performed. The calculated and experimental spectra were in fair agreement for
the lower energy region, including an almost quantitative reproduction of the dye
molar extinction coefficient of the first visible absorption band. According to this
analysis, however, the visible band, at ca. 2.4 eV, was mainly composed by Ligand-
Based Charge Transfer (LBCT), due to a substantial contribution to the HOMOs of
the oxygen lone pairs of the deprotonated carboxylic groups.
Soon after the work by Monat et al., two papers reporting on TDDFT calcula­
tions appeared; one by our group, concerning the N3 complex [13.15], and another
by Guilemoles et al. [13.60], who addressed the ground and excited-state prop­
erties of certain [M(bpy)2L2] complexes (M = Ru, Os, L = CN, SCN, bpy = 2,2'-
bipyridine). The complexes investigated by Guillemoles et al. were reduced models
of N3, in the sense that no carboxylic substituents were present on the 4-4' positions
of the bipyridine ligands. In this study, geometrical optimizations and excited state
calculations were performed both in vacuo and in aqueous solution, and the latter
were modeled by PCM. For the [Ru(bpy)2(NCS)2] complex, a substantial solvation
effect was found: the calculated absorption maxima of the low energy MLCT band
were computed in vacuo (solution) at 554 (482) nm, with oscillator strengths of 0.02
(0.09); the higher energy band was found at 412 (397) nm with oscillator strengths
of 0.08 (0.02). Comparing these results with the experimental solution absorption
maxima at 538 and 398 nm, we notice that the position of the lower energy MLCT
band is considerably affected by the presence of the solvent, shifting towards higher
energies by 0.33 eV with a reversal of the oscillator strengths. However, calcula­
tions in vacuo give rise to a better quantitative agreement with the experimental data
as opposed to calculations in solution. The former gives deviations in the lowest
energy band of 0.07, whereas the corresponding deviation is 0.27 eV for calcula­
tions in solution.
562 Dye-Sensitized Solar Cells

Our work considered the tetra-protonated N3 species [13.15], corresponding


to the experimental species observed at low pH (<1.5), where the four carboxylic
groups are entirely protonated. A geometrical optimization of the N3 complex
showed an excellent agreement between computed and experimental results: the criti­
cal ruthenium-nitrogen bonds were accurately reproduced, as well as the lengthening
of the Ru-N bipyridine bonds trans to the thiocyanate ligands.
A comparison between the electronic structure of the complex, in terms
of frontier molecular orbitals, calculated in vacuo and in solution, is presented in
Figure 13.2.
The HOMO/HOMO-2 are a set of quasi-degenerate orbitals. In agreement with
Rensmo et al. [13.57], the largest orbital contributions arose from the highest occu­
pied thiocyanate orbitals, resulting from a combination of the sulfur p orbitals and
nitrogen lone pairs. These were mixed, in an antibonding fashion, with small percent­
ages of t2g metal orbitals. The HOMO-3, observed 0.25 eV below the HOMO-2, was
localized entirely on the thiocyanate ligands. At lower energy, three quasi-degenerate
orbitals were found, which presented a predominant Ru-d character. The six low­
est LUMOs were the antibonding π* orbitals of the bipyridine ligands, with sizable
contributions from the carboxylic groups (Fig. 13.2). Since, in DSCs, the acidic car­
boxylic units serve as anchoring groups to the Ti0 2 semiconductor surface, such a
contribution from the carboxylic groups to the π* LUMOs, which represent the final
states in MLCT transitions (see below), should favor the electron injection process
from the dye-excited state to the semiconductor surface.
The inclusion of solvation effects led to a sizable change of both energies and
composition of the molecular orbitals. In ethanol, the lowest LUMOs were signifi­
cantly destabilized, by ca. 0.3-0.4 eV with respect to the corresponding values com­
puted in vacuo. The first set of HOMOs was stabilized by 0.3-0.4 eV with respect to
the corresponding values computed in vacuo, and also showed a higher contribution
of ruthenium d orbitals mixed with NCS ligand orbitals. On the whole, the presence of
the solvent caused an important increase in the HOMO-LUMO gap, which changed
from 0.50 eV in the gas-phase to 1.12 eV in ethanol. A similar trend was observed
upon moving from ethanol to a water solution, whereby an even larger HOMO-LUMO
gap of 1.24 eV was found (Fig. 13.2).

Fig. 13.2 Left: The molecular orbital energy for the protonated N3 species in vacuo, in ethanol
and in water solutions, along with the HOMO-LUMO gaps. Right: Isodensity plots of selected
frontier molecular orbitals (L stands for LUMO, H for HOMO). Adapted from Ref. [13.15].
Theoretical and model system calculations 563

The absorption spectra of cis-N3 calculated in vacuo shows two well separated
peaks with intensity maxima at 1.43 and 2.70 eV. The separation of the two bands,
1.27 eV, was significantly larger than the experimental value of 0.8 eV, pointing at a
limited agreement with the experiment. The calculated absorption spectra in ethanol
and water solutions are reported in Figure 13.3.
The overall shape and band separations of the N3 spectrum could be decently
reproduced upon inclusion of solvation effects. This reflected the major changes in the
electronic structure observed in solution. The calculated spectra were in good agree­
ment with their experimental counterparts, apart from a quasi-rigid red-shift of the
entire spectra of ca. 0.3 eV. This inaccuracy was further lifted by using a hybrid XC
functional, such as B3LYP, as can be seen in the right panel of Figure 13.3 where the
comparison between the experimental and B3LYP-calculated absorption spectra for
the tetra-deprotonated N3 complex is reported [13.16,13.17]. The agreement between
theory and experiment was almost perfect throughout an energy range of 4.5 eV. In
addition, these results quantitatively reproduced the solvatochromic shifts exhibited
by the N3 complex with an increasing solvent polarity, which appeared to be related
to a decreased dipole moment in the excited state with respect to the ground state.
As compared to calculations by Monat et al. of the tetra-deprotonated N3 complex
in vacuo [13.58], our TDDFT calculations in solution did not show any signature of
LBCT transitions from the deprotonated carboxylic groups lone pairs. This was due to
the ensuing negative charge, in solution, being effectively screened and these orbitals
being found at lower energies.

13.3.3 Calculations on other Ru(II)-dye sensitizers


In order to improve the DSC efficiency, thousands of new dyes have been scrutinized
and tested. Although delivering record efficiencies, a major issue with N3 and N719
complexes is that they mainly absorb light in the green-blue region of the visible
spectrum, with molar extinction coefficients on the order of 14.000 dm3 mol -1 cm-1.
As a result, they lose out on a considerable amount of red and near-IR light.

Fig. 13.3 Left: Calculated absorption spectra for the protonated N3 complex in ethanol and
water solutions by the BPW91 - xc functional. Right: A comparison between the experimental
and calculated absorption spectra by the B3LYP - xc functional for the tetra-deprotonated N3
complex in water solution. Adapted from [13.15] and [13.17].
564 Dye-Sensitized Solar Cells

Starting from the N3 or N719 homoleptic sensitizers, a major research line has
been devoted to devising heteroleptic sensitizers, in which one of the equivalent two
bipyridines of N3 or N719 is specifically functionalized to give rise to increased
DSC performances. In this context, we briefly mention the results obtained for
two such heteroleptic complexes, i.e., [Ru(4,-carboxylic-acid-4'-carboxylate-2,2'-
bipyridine)(4,4/-di-(2-(3,6-dimethoxyphenyl) ethenyl)-2,2'- bipyridine)(NCS)2], Ν945
[13.19], and [Ru(4,-carboxylic-acid-4/-carboxylate-2,2/-bipyridine)(4,4/-bis[(E)-2-
(3,4-ethylenedioxythien-2-yl)vinyl]-2,2/-bipyridine), Ru-EDOT [13.20], (Fig. 13.4).
In ethanol solution, the Ru-EDOT HOMO was found to be a combination of
Ru t2g and SCN π orbitals, appreciably mixed with π-EDOT character. This is at
variance with the N621 complex, featuring an alky 1-substituted bipyridine ligand, for
which a pure Ru-SCN HOMO, similar to that of N3, was found [13.14]. Such mix­
ing depends on the conjugation with EDOT arms in the bipyridine ligand. Since the
HOMOs are the starting orbitals for the low-lying charge transfer transitions of Ru-
EDOT, vide infra, the partial EDOT contribution to this orbital suggests that, upon
dye photo-oxidation following electron injection into Ti0 2 , the positive charge is
partially localized to the ligand, which becomes stabilized by the electron-donation
from the EDOT substituent. The LUMO is a π* orbital localized on the protonated
carboxy-bipyridine ligand, while the LUMO+1, lying 0.26 eV above the LUMO, is a
7Γ* orbital localized on the EDOT-substituted bipyridine.
The TDDFT absorption spectrum calculated in ethanol (Fig. 13.4) was in good
agreement with the experimental spectrum for the low energy region, whereas a cer­
tain discrepancy in the intensity distribution was found at higher energies. The 538-nm
absorption band could be well reproduced by the theoretical approach (540 nm); this
feature was composed by three major contributions, all having the LUMO+1 as the

Fig. 13.4 Left: Molecular orbital energies and isodensity plots of selected molecular orbitals
for the Ru-EDOT complex. Right: The structure of the Ru-EDOT complex and a compari­
son between its experimental and calculated absorption spectra in ethanol solution. Adapted
from [13.20].
Theoretical and model system calculations 565

arriving state. The lowest excited state originated from the transition to the LUMO,
localized on the carboxylated bipyridine. The calculated absorption spectrum and the
order of excited states was perfectly consistent with the high experimental photo­
voltaic efficiencies. Indeed, excitation at about 540 nm would mainly lead to excited
states localized onto the EDOT-substituted bipyridine. Since these heteroleptic sen-
sitizers were adsorbed onto Ti0 2 via the carboxy-bipyridine ligand, the presence of
lower-lying excited states residing on such ligands permitted an efficient energy trans­
fer to this state, as well as a subsequent electron injection to Ti0 2 . A similar electronic
structure was calculated for the N945 complex, which exhibited an even higher molar
extinction coefficient of 18.900 dm3 mol -1 cm-1, providing an increased photocurrent
when applied in DSCs [13.19].

13.3.4 Trans-complexes
Trans-thiocyanate complexes were also been investigated in order to increase the
dye-sensitizer light-harvesting capability. The electronic structure of the trans
N3 isomer resembled that of the eis isomer discussed previously [13.15]. The
most notable difference was a reduction of 0.18 eV in the HOMO-LUMO gap as
compared to the eis isomer. Such an enhanced red response of the trans isomer
renders trans-thiocyanate complexes attractive candidates for charge-transfer sen-
sitizers in DSCs. In this context, a ruthenium complex trans-[Ru(L)(NCS)2], with

Fig. 13.5 A reflection of representative organic dyes. Upper left: JK1/JK2. Right: D5/D7/D9/
Dll. Bottom left: Squaraine. Adapted from [13.11, 13.12, and 13.13].
566 Dye-Sensitized Solar Cells

L = 4,4///-di-tert-butyl-4/,4//-bis(carboxylic acid)-2,2/:6/,2//:6//,2///-quaterpyridine,
N886, was synthesized by Barolo et al with the aim of enforcing the stability of
the trans thiocyanate coordination [13.18]. Both spectroscopic, electrochemical
and theoretical calculations confirmed an enhanced red-response of this elegantly
designed dye, which showed absorption bands of mixed Ru/SCN-to-quaterpyridine
charge-transfer character extending from the near-IR to the UV regions. However,
despite the favorable localization of the excited states and the red-shifted absorp­
tion spectrum, this complex did not deliver the expected increase in photovoltaic
performances. This was probably due to dye aggregation onto the Ti0 2 surface and
a partially unfavorable alignment of the dye-excited states with the Ti0 2 conduction
band energy.

13.3.5 Organic sensitizers


One major challenge for the engineering of novel sensitizers is to increase their
response in the red part of the near-IR region of the solar spectrum and to get them
to display an increased molar extinction coefficient. While extended light-harvesting
is generally required for efficient sensitizers in both liquid and solid-state DSCs, the
demand for highly absorbing dyes becomes dramatic in solid state devices, where
a semiconductor film of reduced thickness is employed for maximum efficiency. In
this respect, organic dye molecules are attractive materials due to their potentially
high molar extinction coefficient, adjustable absorption spectral response, and envi­
ronmental compatibility. As an example, the widely investigated indolene dyes show
broad absorptions at ca. 550 nm with molar extinction coefficients on the order of
40.000 dm3 mol -1 cm -1 [13.10], while squaraine dyes display tuneable far-read or near-
IR absorptions with molar extinction coefficients exceeding 150.000 dm3 mol -1 cm -1
[13.12]. Such properties render them promising alternatives to Ru(II) dyes as light-
absorbing molecules, especially in thin-film DSCs.
Several classes of full organic sensitizers have been developed, see Fig. 13.5,
including the aforementioned indolene [13.10] and squaraine [13.12] dyes, as well
as novel Donor-7r-Acceptor (D-π-Α) architectures [13.11, 13.13]. Here, we focus
on a few selected representative theoretical calculations on the organic dyes reported
in Figure 13.5, namely the JK1/JK2 dyes [13.11], a squaraine dye [13.12] and the
so-called D5/D7/D9/D11 dye series [13.13]. JK1/JK2 and the D5/D11 series are pro­
totypical dyes designed according to a D-π-Α architecture, characterized by a donor
moiety (a tertiary aminé), a spacer (thiophene) and an acceptor moiety (cyanoacrylic
acid). In the squaraine case, an asymmetry is introduced in the molecule to provide
vectorial charge transfer towards the carboxylic function.
For JK1-JK2 and the D-series, the HOMOs are, as expected, generally localized
on the donor portion of the molecule while the LUMOs are present on the cyanoacrylic
acceptor moiety (Fig. 13.6). Also a substantial conjugation across the thiophene
bridges can be observed in both cases. It is interesting to notice that the HOMO and
LUMO of JK-2 are destabilized and stabilized, respectively, as compared to JK-1,
thus reducing the HOMO-LUMO gap from 2.47 to 2.21 eV.
TDDFT calculations performed on JK1 revealed a lowest transition at 2.23 eV,
corresponding to a charge-transfer (CT) excitation from the bis-dimethylfluoreneanili
Next Page

Theoretical and model system calculations 567

Fig. 13.6 Left: Isodensity plots of selected frontier molecular orbitals for JK1. Right: A sche­
matic representation of the energy levels for JK1 and JK2 in vacuo and in ethanol solution.
Images for thisfigurekindly supplied by Dr. A. Vittadini.

ne-based HOMO to the LUMO, localized on the thiophene-cyanoacrylate moieties. As


compared to the experimental absorption maximum, found at 2.84 eV, the calculated
transition was considerably red-shifted. This was related to the extended charge-trans­
fer character of this transition, which was not properly captured by TDDFT calculations
employing current xc functionals [13.48]. A better agreement between the calculated
and experimental absorption energies of the higher-lying π-π* features as com­
pared to the CT excitation was due to the localized character of the π-π* excitations,
involving substantially overlapping orbitals. The calculated molecular orbitals of D5,
D9, and D l l were rather similar to those of JK1-JK2. The lowest TDDFT excitation
energies calculated for these systems, although qualitatively reproducing the similar
spectroscopic behavior of D9 and D l l , considerably underestimated both the absorp­
tion maxima and the estimated E(0_0) transitions. This corresponded well to what was
found for JK-1 and JK-2. Improved xc functionals are currently under development
and should allow us to overcome the difficulties with regard to the description of
charge-transfer-excited states in organic dye sensitizers.

13.3.6 Squaraine dyes


Essential for an efficient conversion of solar energy by DSCs is the spectral match
of the sensitizer absorption to the solar radiation. In this regard, squaraines are well-
known for their intense absorption in the red/near-IR regions. A novel asymmetrical
squaraine sensitizer with a carboxylic acid group directly attached to the chromo-
phore has been synthesized [13.12]. The lowest excited state is calculated to give rise
to an isolated intense transition at 596 nm (/= 1.301), i.e., only blue-shifted 0.13 eV
as compared to the experimental absorption maximum. The resulting excited state is
potentially strongly coupled to the semiconductor surface, due to charge delocaliza-
tion involving the anchoring carboxylic group. This factor, together with the high dye
INDEX

Index Terms Links

accelerated long-term tests 273


activated carbon electrode 30
active iodide molten salts 132
additives 123
in ILEs 139
effect on DSC performance 166
admittance spectroscopy 2
adsorption and excited states 579
aerosol pyrolysis 59 63
alizarin-TiO2 interface 578
alternative redox mediator electrolytes 447
ambipolar diffusion coefficient 122
amorphous organic semicondutor 175
amorphous titania 59
amphiphilic heteroleptic complex 92
anastate, crystal structure 47
anatase 23 45 47
nanotubes 69
electrochemical characteristics 50
anchoring groups 28 86
angle-of-incidence effect 282
anodic oxidative hydrolysis 62
artificial photosynthesis 2

This page has been reformatted by Knovel to provide easier navigation.


Index Terms Links

artistically designed DSCs 285


arylamine electron donors 27

back electron transfer 119 124


back electron transfer rate 432
back-contact DSCs 294
ball-grid DSC 302
band bending 11
band-edge shift 96
band-gap energy for rutile 47
bilayer structure of the TiO2 electrode 300
bimodal films 74
bimolecular quenching 152
binary melts 138
black dye 293
block copolymers 198
blocking layers 356
bode phase plots 95 366
bridging thiophene 103
brookite 23 45
buffering effect of traps 329
building-integrated photovoltaics 28 35 269 277

C60 cluster film 149



C60/C60 redox couple 149

This page has been reformatted by Knovel to provide easier navigation


Index Terms Links

cadmium chalcogenides 13
calculation of diffusion lengths 183
camouflaged design 316
capacitance of electron accumulation 32
carbon black 30
carbon black-based counter electrodes 23
carbon nanotubes 70
carrier trapping 174
cation effect 122
cations 121
cell-to-cell interconnects 10
leakage 22
chalcogenide 13
charge accumulation mechanism 56
charge carrier mobility 175 176
charge diffusion transport 135
charge dissociation 121
charge extraction methods 377
charge injection 153
dynamics 416
efficiency 21
charge recombination 192 194 430
charge rectification effect 147
charge separation efficiency 148
charge separation of organic clusters 154
charge transfer
processes 118 150
reactions 406
resistance 355

This page has been reformatted by Knovel to provide easier navigation


Index Terms Links

charge transport 174


in solid-state DSCs 185
charge-carrier diffusion length 15
generation 8
chemical capacitance 70 364
p-doping 177
chlorophyll 16
analogues 145
Co(II)/Co(III) 29
colloids 14
color solar-cell design 297
colored DSC modules 311
ColorSol 277
commercial applications 34
complex impedance plot 17
composite semiconductor films 147
conductive plastic substrates 31
conductivity of spiro-MeOTAD 180
conjugation bridge 92
constant phase element 21
consumer electronics applications 8
products 37
continuity equation 333
controlled solvolysis of TiCl4 65
conversion efficiency 20
co-sensitization 113
counter electrode 29
performance 22
counter-Pt electrodes 258
courmarine dye 102

This page has been reformatted by Knovel to provide easier navigation


Index Terms Links

croconate dyes 154


current collection 181
cyclic voltammogram 344

Dahms-Ruff equation 120


Density Functional Theory 557
depletion regime of TiO2 55
design 270
determining injection rates experimentally 420
de-trapping 174
device degradation 272
DFT calculations 101
dichalcogenides 17
dicyanoamide 135
diffusion 47
and drift 48
coefficient 53
coefficients of electrons 120
length 183
model of electron generation 339
of the triiodide ion 120
diffusion-recombination model 2 63
diffusion-recombination transmission line 59
dimethylfluoreneaniline moieties 105
diode equation 7
dipolar aprotic solvents 121
directionality 84
dithienothiophene unit 106

This page has been reformatted by Knovel to provide easier navigation


Index Terms Links

donor-p-acceptor 566
doping of TiO2 63
“double gyroid” diblock copolymer phase 200
double-layer electrode 75
double-layer charging 56
double-layer film 75
down-converters 11
drop-in hole-transfer efficiency 191
DSC
assembling 258
electrolytes 28
solar steel roofing panels 37
substrates 22
efficiency parameters 19
monolithic design 31
performance factors 323
p-type semiconductor 37
quasi-solid state 18
remote industrial applications 38
solid state 163
Spiro-OMeTAD based 165
based on ionic liquids 171
incorporating CuSCN 164
with plastic substrates 290
dye regeneration kinetics 172
dye regeneration mechanism 173
dye sensitization 16 414
dyes for DSC 24

This page has been reformatted by Knovel to provide easier navigation


Index Terms Links

edge-sharing TiO6 octahedra 69


effect of alkyl chains 89
effective electron lifetime 330
electrocatalyst 29
electrochemical deposition of titania 62
electrochemical etching 153
electrochemical impedance spectroscopy 1 95
electrochemically deposited
nanocrystalline ZnO 299
electrochemistry of titania 55
electrodeposition 30
electrodeposition of TiO2 71
electroinactive imidazolium salts 135
electrolyte components 121
electrolyte of low viscosity 132
electron-back-transfer reaction resistance 20
electron
conductor 77
diffusion length 183 386
donating alkoxy groups 96
exchange diffusion coefficient 120
hole recombination 184 194
injection 62
kinetics 119
lifetimes of the complex 95
mediators 125
relay 149
transfer 18 168

This page has been reformatted by Knovel to provide easier navigation


Index Terms Links

electron (Cont.)
transfer kinetics 146
transfer to I3* 354
transport in nanocrystalline TiO2 films 450
transport in the DSC 331
trapping 328
electron-acceptors (p-doping) 8
electron-donors (n-doping) 8
electron-hole pairs 49
electrosynthesis of tubular pores 71
embodied energy for DSC 3
emission lifetime 152
energetic disorder 174
energetics of the oxide/dye/
electrolyte interface 342
energy gap dependence 146
energy level diagram 150 342
energy levels at the heterojunction 173
enhanced conductivity 181
enhancement in mobility 181
enhancing light capture 195
equilibrium electron density 326
equivalent circuits 14
ETA (extremely thin absorber) solar cell 36
excited molecules, energy levels 349
excited state oxidation potential 572
excited state sensitizer 151
excited-state lifetime of Ru(bpy)32+ 146

This page has been reformatted by Knovel to provide easier navigation


Index Terms Links

fabrication methods 251


facade applications 276
faradaic process 62
ferrocene/ferrocenium 126
Fick’s second law 335
fill factor 84
film-forming properties 178
first-generation devices 9
flatband potential 72 121 344
foamed-spray pyrolysis 306
fractal anatase layers 62
fractal aspects of mesoporous films 24
fractal dimension 24
frequency response analysis 362
Freundlich’s analysis 4

gelators 140
gelled electrolytes 139
general diffusion transmission lines 56
Gerischer diagram 352
glass-based dye cells 280
glass-facade elements 277
greenhouse gases 4
grid-connected solar farms 37
ground state oxidation potential 572
gyroid network 63
This page has been reformatted by Knovel to provide easier navigation
Index Terms Links

haze 74
heptylimidazolium iodide 311
hermetic seals 18
heterocyclic compounds 123
heterogeneous photocatalaysis 15
heteroleptic complexes 99
high hole mobility 175
high molar extinction coefficient sensitizers 96
highly dispersive charge transport 175
high-pressure titania polymorphs 48
high-surface-area mesoporous electrodes 312
hole conductors 36 352
hole-mobility 181
hole-transfer 171
hole-transport material (HTM) 32
hole-transporter infiltration 189
hole-transporting media 448
HOMO-LUMO gap 571
homolytic fission 2
hopping model 174
hot carrier cell 11
hot electron injection 147
hot-melt encapsulation 304
H-type and J-type aggregates 154
HX decomposition 5
hydrolysis of alkoxides 62
hydrolysis of titanium(IV) alkoxides 61
hydroquinone 17

This page has been reformatted by Knovel to provide easier navigation


Index Terms Links

hydrothermal growth 61
hydrothermal methods 202
hydrothermal synthesis 28

ideal diode model 5


IiO2, orthorhombic 45
imidazolium-based ionic liquids 275
imidazolium iodides 133 138
impedance model 24
impedance of the cathode 366
impedance spectroscopy 1 185 363
IMVS response 371
incident photon-to-current
conversion efficiency (IPCE) 49 85 336 338
indium-doped tin oxide 22
indolene-based organic sensitizer 171
indoline dyes 306
indoor applications 305
indoor electronics 295
influence of ionic salts 180
injection mechanism 586
inorganic nanoparticles 141
in-situ FTIR spectroelectrochemistry 52
in-situ Raman spectroelectrochemistry 60
intensity modulated photocurrent
spectroscopy 183 369
interband transitions 59
intercalation 57

This page has been reformatted by Knovel to provide easier navigation


Index Terms Links

interfacial electron transfer 324


intermolecular quenching 104
internal current collectors 10
intraband trap states 411
inverse opal 63 67
iodide/triiodide couple 28
iodide-based electrolytes 441
ion solvation 193
ionic liquid electrolyte 128

key components of the DSC 21


kinetic treatments 325
Kohlrausch-Williams-Watts (KWW)
function 390
Kohn-Sham (KS) equations 557
Kubelka-Munk 337

lamp prototype 295


large-area prototype 292
latent image 5
lateral hole-hopping 352
+
Li adsorption on TiO2 123
light soaking 269
light-harvesting efficiency 54 336
light-scattering TiO2 layer 262
light-weight flexible DSC panels 12
This page has been reformatted by Knovel to provide easier navigation
Index Terms Links

limiting current density 354


liquid electrolyte 117
liquid electrolyte cells 67
liquid-junction photoelectrochemical cells 6 12
Li-storage in nanocrystals 58
lithium salts 170 180
Li-TSFI 181
long-term stability 18 19 127 135
low-viscosity solvents 29
low-volatile electrolyte 135
luminescence 16

manufacturing 31
Marcus theory 4
market segments for PV 34
mass transport of the iodide and triiodide 120
masterplate design 270
material requirements 31
M-design 8
merocyanine orange dye 285
mesoporous films 65
mesoporous TiO2 67
mesoscopic metal oxide semiconductors 406
mesostructured materials, one-pot route 199
metal-free organic sensitizers 102
metal-to-ligand charge transfer transitions 96
methoxypropionitrile 87
military applications 316

This page has been reformatted by Knovel to provide easier navigation


Index Terms Links

modeling of combined dye/semiconductor


systems 556
modules 5 22 267
encapsulation 275
performance 12
flexible plastic-type DSCs 292
molecular aggregation on the semiconductor 103
molecular energy storage reactions 3
molecular structure of the sensitizer 197
molten iodide salts 133
monoclinic 45
pentatitanate 69
monodispersed colloids 15
monolithic design 8 275 281 302
monolithic series-interconnected module 284
Monte Carlo methods 332
morphology of the hole-transporter 178
mosaic nano-arrays 63
Mott-Schottky plot 2 49
+
multi-adsorption of DMHI 123
multimodal structures 74
multiple excitonic charge-carrier generation 11
multiple trapping 49 51 174

N3 complex 560
N719 isomers 587
nanocrystals 45
nanofibrous TiO2 59 69

This page has been reformatted by Knovel to provide easier navigation


Index Terms Links

nanosecond laser-flash photolysis 391


nanosecond time scale 153
nanosheet 59
nanostructuring 33
nanotubes 84
devices 35
near infrared transmittance 381
near-IR absorbing sensitizers 109 313
negative movement 124
nitrogen-containing heterocyclic
compounds 124
N-methoxypriopionitrile MPN 28
nonactive iodide molten salts 135
nonadiabatic molecular dynamics 559
non-glass/TCO substrates 29
non-ideality 327
non-organized titania 61
non-thiocyanato ruthenium complexes 101
norbornadiene 3
n-type doping 47

off-grid rural areas 280


open-circuit potential 84
optimizing cell geometry 15
ordered arrays 153
organic electrolyte 17
organic gelator 139

This page has been reformatted by Knovel to provide easier navigation


Index Terms Links

organic molecular and polymer


semiconductors 164
organic sensitizers 27 102 196
organic solvents 121
organized mesoporous underlayer 75
organized nanocrystalline titania 64
organized TiO2 69
organized TiO2 materials 59
orthorhombic lepidocrocite 69
Ostwald ripening 61
outdoor installation 288
oxazines 145

P-25 powder 74
panchromatic 84
panels 5
temperature dependence 2
parallel connected modules 6
particulate systems 15
π-conjugation 196
p-CuInSe2 37
P-design 6
PEN (polyethylene naphthalate) 290
peptization of the TiO2 64
performance per active area 15
peryleneimide sensitizer 37
perylenes 145
photocorrosion 13 14

This page has been reformatted by Knovel to provide easier navigation


Index Terms Links

photocurrent decay measurements 362


photocurrent density 49
photoelectrochemistry 49
photogalvanic cell 6
photogalvanic mechanism 149
photoinduced absorption 190
photoinduced absorption spectroscopy 388
photo-induced electron transfer process 168
photolysis of thionine 7
photonic crystal 67
photo-redox reactions 4 63
photosensitization efficiency 146
photosensitizer 3
photosynthesis 2
Photovoltage decay 374
photovoltaic
action spectra 191
performances 86
production 1
smart windows 312
phthalocyanines 109
plastic DSC modules 290
platinum-catalyst 30
Pluronic-123 198
Pluronic-templated TiO2 65
p-NiO 37
polaron formation 174
polaron-free system 174
poly(3-octylthiopehene) 33
polyaniline 33

This page has been reformatted by Knovel to provide easier navigation


Index Terms Links

polycrystalline anatase 51
polycrystalline materials 13
polycrystalline silicon panels 1
polydiacetylenes 33
polymers with a conductive coating 29
polypyridyl cobalt (II/III) complex 126
polystyrene opal 68
porosity of mesoscopic oxide films 145
porous titania 60
porous-TiO2 electrodes 254
portable electronics 267
portable power packs 278
potential-step chronoamperometry 58
power source for unmanned aerial vehicles 317
producing titania electrodes 73
product lifetime 18
Prussian blue 312
pseudocapacitive Li-storage in TiO2 57
p-type organic semiconductors 33
pump-probe spectroscopy 148
pyridine derivatives 124

Q-size effects 63
quantum confinement effect 51
quantum yield of charge injection 54
quarternary ammonium iodides 139
quasi Fermi 326 383

This page has been reformatted by Knovel to provide easier navigation


Index Terms Links

Quasi solid-state
electrolytes 117 139 140 275
DSCs 32
quasi-static approximation 333
quenching rate constant 151
quenching, oxidative and reductive 4

recombination 19
kinetics 390
reaction 119
resistance 35 73 90
redox mediators 17 28 440
redox systems, energy levels in solution 348
reductive quenching 171
regenerative solar cell 12
remote-community applications 39
reproducibility of DSC photovoltaics 264
resonance raman spectroscopy 147
rhodamine 145
role of trap states 438
roll-to-roll manufacturing 278
rooftop DSC modules 287
room-temperature ionic-liquid electrolytes 128
room-temperature ionic liquids 18
roughness factor 24 74
Ru(bpy)32+ 52

This page has been reformatted by Knovel to provide easier navigation


Index Terms Links

ruthenium
complexes 83 99
sensitizers 84
demand for the electronics industry 26
rutile 23 45
crystal structure 47

screen photo-injected electrons 121


screen-print procedure 67 255
(SeCN)33–/SeCN– couple 127
seals 9
second generation devices 9
see-through type DSCs 304
self-organization of anatase 64
self-organized TiO2 34
self-quenching 7
self-trapping 174
semiconductor electrodes 11
semiconductor oxide substrates 145
semiconductor-liquid interface 408
sensitization of nanotube arrays 153
sensitized photocurrent generation 154
sensitizer 83 84
series resistance 22 89
shielding 46
shift of the TiO2 conduction band 169
Shockley and Quiesser, calculation of 10
Shockley-Queisser limit 25

This page has been reformatted by Knovel to provide easier navigation


Index Terms Links

short-circuit photocurrent 84
SILAR 36
simulated device performance 14
single strip cells 8
single-crystal anatase electrode 71
single-junction photovoltaic cell 99
sintering of nanoparticles 73
small polaron 174
small-area solar cells 267
SnO2 145
SnO2-C60-Ru(bpy)32+ 149
(SCC) solar-energy chemical converter 14
solar array 267
solar energy 1
solar panels 267
solar-cell installation in clothing 307
solar-powered functional traffic signs 307
sol-gel procedure 55
solidification 141
solid-state DSC, key components 164
solid-state electrolyte 117 163
solvation effects 558
solvent-free ionic-liquid electrolyte 107
space-charge layer 55
spectral properties of ruthenium sensitizers 84
spectral sensitization of TiO2 52
Spectroelectrochemical measurements 147
Spectroelectrochemistry of titania 59
spiro-OMeTAD 32 164
spray-pyrolysis-deposition method 287

This page has been reformatted by Knovel to provide easier navigation


Index Terms Links

squaraine 110 145 154 567


stainless-steel-based TiO2-film electrode 309
sub-band-gap excitation 52
substrates and superstrates 28
sun 1
supramolecular templating 70
surface potential of the TiO2 169
surface-trap-mediated recombination 341
surfactant-templated mesoporous anatase 59
surfactant-templating 64
SWCNT network 153

TDDFT excitation energies 567


temperature-humidity cycling test 269
templating by polystyrene spheres 67
templating with block copolymers 65
ter-butylpyridine 123
tetraalkyl phosphonium iodides 139
tetragonal TiO2 46
theoretical calculations 556
thermal decomposition of Ti(OR)4 62
thermal-cycling test 269
thermalization losses 11
thermoplastic gaskets 10
thin-film solar panels 269
thiocyanate 135
thiophene alkyl-chain 104
Third generation solar cells 11 25

This page has been reformatted by Knovel to provide easier navigation


Index Terms Links

TiCl4 treatments 260


time-resolved photoluminescence decay 171
TiO2
conduction band 109
crystalline forms 23
models 568
nanotubes 85
pastes 252
pore filling 187
monoclinic 45
titanate nanowires 70
titania 45
filler 141
nanotubes 69 70
polymorphs 45
depletion regime 48
Toyota Dream House 285
transient absorbance signals 405
transient absorption decay kinetic study 127
transient absorption spectroscopy 147 193
transient photocurrent and photovoltage
decay method 183
transient photocurrent data 123 184
transient photoinduced absorption
spectroscopy 172
transient photovoltage 273
transmission-line models 52
transparent conducting oxide 294
transparent counter-electrodes 285
transparent mesoporous TiO2 electrodes 344

This page has been reformatted by Knovel to provide easier navigation


Index Terms Links

transport resistance 44 77
trans-thiocyanate complexes 565
trap-controlled transport 174
trap-limited devices 360
trapping/detrapping effects 372
triiodide concentration 125
triiodide/iodide redox couple 126
tunability of the scattering properties 293
tungsten interdigitated electrodes 301
tuning spectral response 99

ultrafast electron injection 119 403


universal solar charger 278
up-converters 11

viscosity behavior of ionic liquids 132


voided diblock copolymer film 200

water oxidation 63
water photolysis 5
W-design 7 310
white color DSCs 300
W-interconnected modules 7

This page has been reformatted by Knovel to provide easier navigation


Index Terms Links

WO3 layer 312


World Solar Car Rally 305

xanthene dyes 7

Z-design 7 270 310


ZnO 145
nanowires 34
Zn-prophyrin sensitizers 196
Zr-doped anatase 62

This page has been reformatted by Knovel to provide easier navigation

Das könnte Ihnen auch gefallen