0% found this document useful (0 votes)
2K views284 pages

Solar Energy Conversion Storage: Photochemical Modes

Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
0% found this document useful (0 votes)
2K views284 pages

Solar Energy Conversion Storage: Photochemical Modes

Copyright
© © All Rights Reserved
We take content rights seriously. If you suspect this is your content, claim it here.
Available Formats
Download as PDF, TXT or read online on Scribd
You are on page 1/ 284

SOLAR ENERGY

CONVERSION
and STORAGE
Photochemical Modes
ELECTROCHEMICAL ENERGY STORAGE AND CONVERSION
Series Editor: Jiujun Zhang
National Research Council Institute for Fuel Cell Innovation
Vancouver, British Columbia, Canada

Published Titles

Electrochemical Supercapacitors for Energy Storage and Delivery: Fundamentals and Applications
Aiping Yu, Victor Chabot, and Jiujun Zhang
Proton Exchange Membrane Fuel Cells
Zhigang Qi
Graphene: Energy Storage and Conversion Applications
Zhaoping Liu and Xufeng Zhou
Electrochemical Polymer Electrolyte Membranes
Yan-Jie Wang, David P. Wilkinson, and Jiujun Zhang
Lithium-Ion Batteries: Fundamentals and Applications
Yuping Wu
Lead-Acid Battery Technologies: Fundamentals, Materials, and Applications
Joey Jung, Lei Zhang, and Jiujun Zhang
Solar Energy Conversion and Storage: Photochemical Modes
Suresh C. Ameta and Rakshit Ameta

Forthcoming Titles
Electrochemical Energy: Advanced Materials and Technologies
Pei Kang Shen, Chao-Yang Wang, San Ping Jiang, Xueliang Sun, and Jiujun Zhang
Solid Oxide Fuel Cells: From Fundamental Principles to Complete Systems
Radenka Maric
ELECTROCHEMICAL ENERGY STORAGE AND CONVERSION

SOLAR ENERGY
CONVERSION
and STORAGE
Photochemical Modes

Edited by
Suresh C. Ameta
Rakshit Ameta
CRC Press
Taylor & Francis Group
6000 Broken Sound Parkway NW, Suite 300
Boca Raton, FL 33487-2742
© 2016 by Taylor & Francis Group, LLC
CRC Press is an imprint of Taylor & Francis Group, an Informa business

No claim to original U.S. Government works


Version Date: 20150916

International Standard Book Number-13: 978-1-4822-4631-5 (eBook - PDF)

This book contains information obtained from authentic and highly regarded sources. Reasonable efforts have been
made to publish reliable data and information, but the author and publisher cannot assume responsibility for the valid-
ity of all materials or the consequences of their use. The authors and publishers have attempted to trace the copyright
holders of all material reproduced in this publication and apologize to copyright holders if permission to publish in this
form has not been obtained. If any copyright material has not been acknowledged please write and let us know so we may
rectify in any future reprint.

Except as permitted under U.S. Copyright Law, no part of this book may be reprinted, reproduced, transmitted, or uti-
lized in any form by any electronic, mechanical, or other means, now known or hereafter invented, including photocopy-
ing, microfilming, and recording, or in any information storage or retrieval system, without written permission from the
publishers.

For permission to photocopy or use material electronically from this work, please access www.copyright.com (http://
www.copyright.com/) or contact the Copyright Clearance Center, Inc. (CCC), 222 Rosewood Drive, Danvers, MA 01923,
978-750-8400. CCC is a not-for-profit organization that provides licenses and registration for a variety of users. For
organizations that have been granted a photocopy license by the CCC, a separate system of payment has been arranged.

Trademark Notice: Product or corporate names may be trademarks or registered trademarks, and are used only for
identification and explanation without intent to infringe.
Visit the Taylor & Francis Web site at
http://www.taylorandfrancis.com
and the CRC Press Web site at
http://www.crcpress.com
Contents
Preface..............................................................................................................................................vii
Editors................................................................................................................................................ix
Contributors.......................................................................................................................................xi

Chapter 1 Introduction................................................................................................................... 1
Suresh C. Ameta

Chapter 2 Photochemical Solar Energy Conversion......................................................................7


Rakshit Ameta, Chetna Ameta, and Poonam Kumawat

Chapter 3 Basic Photoelectrochemistry....................................................................................... 17


Purnima Dashora, Meenakshi Joshi, and Suresh C. Ameta

Chapter 4 Photoelectrochemical Cells......................................................................................... 29


Dipti Soni, Priya Parsoya, Basant K. Menariya, Ritu Vyas, and Rakshit Ameta

Chapter 5 Organic Photovoltaic Cells.......................................................................................... 55


Meenakshi Singh Solanki, Taruna Dangi, Paras Tak, Sanyogita Sharma, and
Rakshit Ameta

Chapter 6 Dye-Sensitized Solar Cells.......................................................................................... 85


Rakshit Ameta, Surbhi Benjamin, Shweta Sharma, and Monika Trivedi

Chapter 7 Photogalvanic Cells................................................................................................... 115


Yasmin, Abhilasha Jain, Pinki B. Punjabi, and Suresh C. Ameta

Chapter 8 Hydrogen: An Alternative Fuel................................................................................. 139


Neelu Chouhan, Rajesh Kumar Meena, and Ru-Shi Liu

Chapter 9 Photocatalytic Reduction of Carbon Dioxide............................................................ 173


Guoqing Guan, Xiaogang Hao, and Abuliti Abudula

Chapter 10 Artificial Photosynthesis........................................................................................... 187


Neelam Kunwar, Sanyogita Sharma, Surbhi Benjamin, and Dmitry Polyansky

Chapter 11 Nanomaterials for Solar Energy................................................................................ 219


Mohammad Azad Malik, Sajid Nawaz Malik, and Asma Alenad

Chapter 12 Other Solar Cells....................................................................................................... 253


Rakshit Ameta

Index............................................................................................................................................... 265

v
Preface
Energy is a fundamental requirement for society. It is essential for industrialization, transportation,
urbanization, food materials, and so on. The main conventional energy sources are wood,
coal, petrol, diesel, kerosene, and so forth. These sources are being depleted at an ever-increasing
pace, resulting in an era of energy crisis. Natural resources are limited, and it has been estimated
that they will be completely exhausted in the coming few decades. Therefore, there is an urgent
need to find some alternate energy sources to fulfill the energy demands of the world. In this con-
text, researchers are stressing the use of solar energy, because it is abundant, inexhaustible, eco-
friendly, and relatively low cost. Solar energy can be used to generate electricity or can be stored as
chemical energy in the form of hydrogen or reduced products of carbon dioxide.
Various methods for converting light energy into electrical energy involve the use of photoelec-
trochemical cells, dye-sensitized solar cells, organic photovoltaic cells, photogalvanic cells, and
so on. Every method has its own merits or demerits, which can be overcome by continuous and
dedicated efforts. Some newer solar cells have been developed by combining or modifying these
existing cells. Some examples are plasmonic solar cells, hybrid solar cells, biohybrid solar cells,
perovskite solar cells, tandem solar cells, inverted tandem solar cells, and so on. Some of these cells
have shown promise as future sources of energy.
Hydrogen can be obtained by photosplitting water, which is abundantly available on Earth.
Hydrogen has also been advocated as the fuel of the future, because it is nontoxic and has a higher
storage capacity. It generates electricity on burning in a fuel cell. The reduction of carbon dioxide
to some useful synthetic fuels is another possible system for storing solar energy. Nature does this
job by reducing carbon dioxide from the atmosphere in the form of biomass (carbohydrates) and
regenerating oxygen. This conversion is the well-known process of photosynthesis. Efforts are being
made to mimic this reaction in laboratory conditions (i.e., artificial photosynthesis). Although natu-
ral photosynthesis seems to be a simple chemical process, it is mechanistically complex. This is a
challenge for chemists and biologists, in general, and photochemists in particular.
The development of nanotechnology has added many new frontiers with varied applications.
Nanomaterials have smaller size, higher surface-to-volume ratio, and so on, and as a result, they
may have some astonishing properties. In the last decade, nanoparticles have been extensively used
in different forms (nanorods, nanowire, nanoribbon, etc.), and the time is not far off when these
materials will prove their importance in solar cells, enhancing their efficiency as well as reducing
water, carbon dioxide, and other materials.
In this book, the focus is on photochemical methods of converting and/or storing light energy
in the form of electrical or chemical energy. Although efforts have been made to incorporate major
work done in this field, due to certain limitations, some important work may have been left out.
Readers are welcome to suggest any further improvement in this effort.

Suresh C. Ameta
Rakshit Ameta

vii
Editors
Suresh C. Ameta obtained his master’s degree from the University of Udaipur, India, and was
awarded a Gold Medal in 1970. He secured a first position in master of philosophy in 1978 in
Vikram University, Ujjain (Madhya Pradesh, India). He also obtained a PhD from this univer-
sity in 1980. He served as professor and head of the Department of Chemistry, North Gujarat
University Patan, India (1994) and M. L. Sukhadia University, Udaipur (2002–2005), and served
as head of the Department of Polymer Science (2005–2008). He also served as dean of postgradu-
ate studies, M. L. Sukhadia University, Udaipur (2004–2008). Now, he is serving as dean of the
Faculty of Science, PAHER University, Udaipur. Professor Ameta has occupied the coveted posi-
tion of president, Indian Chemical Society, Kolkata, and is now a lifelong vice president (since
2002). He has been the recipient of a number of prizes during his career, including a national prize
(twice) for writing chemistry books in Hindi, the Professor M. N. Desai Award, the Professor W.
U. Malik Award, the National Teacher Award, the Professor G. V. Bakore Award, and a Lifetime
Achievement Award from the Indian Chemical Society. He has successfully guided 71 doctoral
students. Professor Ameta has more than 300 research publications to his credit in national and
international journals and has served as a reviewer. He has contributed to and authored about 40
undergraduate and postgraduate books (published by Nova Publishers, New York; Taylor & Francis
Group, Oxford, United Kingdom; and Trans-Tech Publications, Pfaffikon, Switzerland), including
books on green chemistry and microwave-assisted organic synthesis (published by Apple Academic
Press, Waretown, New Jersey). In addition, he has delivered lectures and chaired sessions at vari-
ous conferences. He completed five major research projects from different funding agencies, such
as the Department of Science and Technology, the University Grants Commission, the Council of
Scientific and Industrial Research, and the Ministry of Energy, Government of India. Professor
Ameta has approximately 43 years of experience in teaching and research.

Rakshit Ameta obtained first position, master of science degree and was awarded a Gold Medal in
2002. He received the Fateh Singh Award from the Maharana Mewar Foundation, Udaipur, India,
for his meritorious performance. He obtained a PhD in 2005 from M. L. Sukhadia University,
Udaipur, India. He has worked at that university as well as at the University of Kota, Kota, India,
and presently is an associate professor of chemistry at PAHER University, Udaipur. He has suc-
cessfully supervised five doctoral students, and seven more are now researching various aspects of
green chemistry. He has authored 70 research publications in national and international journals.
Dr. Ameta has organized many national conferences, delivered a number of invited lectures, and
chaired sessions at national conferences. He has been elected as scientist in charge, Industrial and
Applied Chemistry Section, Indian Chemical Society, Kolkata (2014–2016), and was also elected
as a council member of the Indian Chemical Society, Kolkata (2011–2013), and the Indian Council
of Chemists, Agra (2012–2014). He has authored five degree-level books and contributed chap-
ters in books (published by Nova Publishers, New York; Taylor & Francis Group, Oxford, United
Kingdom; and Trans-Tech Publications, Pfaffikon, Switzerland). Two books have been published
on the topics of green chemistry and microwave-assisted organic synthesis (Apple Academic Press,
Waretown, New Jersey).

ix
Contributors
Abuliti Abudula Purnima Dashora
North Japan Research Institute for Department of Chemistry
Sustainable Energy PAHER University
Hirosaki University Udaipur, India
Aomori City
and Guoqing Guan
Graduate School of Science and Technology North Japan Research Institute for Sustainable
Hirosaki University Energy
Hirosaki, Aomori, Japan Hirosaki University
Aomori City
Asma Alenad and
School of Materials and Chemistry Graduate School of Science and Technology
University of Manchester Hirosaki University
Manchester, United Kingdom Hirosaki, Aomori, Japan

Chetna Ameta Xiaogang Hao


Department of Chemistry School of Chemistry and Chemical
PAHER University Engineering
Udaipur, India Taiyuan University of Technology
Taiyuan, People’s Republic of China
Rakshit Ameta
Abhilasha Jain
Department of Chemistry
Department of Chemistry
PAHER University
St. Xavier’s College
Udaipur, India
Mumbai, India
Suresh C. Ameta Meenakshi Joshi
Department of Chemistry Department of Chemistry
PAHER University PAHER University
Udaipur, India Udaipur, India

Surbhi Benjamin Poonam Kumawat


Department of Chemistry Department of Chemistry
PAHER University PAHER University
Udaipur, India Udaipur, India

Neelu Chouhan Neelam Kunwar


Department of Pure and Applied Chemistry Department of Chemistry
University of Kota PAHER University
Kota, India Udaipur, India

Taruna Dangi Ru-Shi Liu


Department of Chemistry Department of Chemistry
PAHER University National Taiwan University
Udaipur, India Taipei, Taiwan, Republic of China

xi
xii Contributors

Mohammad Azad Malik Sanyogita Sharma


School of Materials and Chemistry Department of Chemistry
University of Manchester Pacific Institute of Technology
Manchester, United Kingdom Udaipur, India

Shweta Sharma
Sajid Nawaz Malik
Department of Chemistry
School of Chemical and Materials
PAHER University
Engineering
Udaipur, India
National University of Sciences and
Technology
Meenakshi Singh Solanki
Islamabad, Pakistan
Department of Chemistry
PAHER University
Rajesh Kumar Meena Udaipur, India
Department of Pure and Applied Chemistry
University of Kota Dipti Soni
Kota, India Department of Chemistry
PAHER University
Udaipur, India
Basant K. Menariya
Department of Chemistry
Paras Tak
PAHER University
Department of Chemistry
Udaipur, India
PAHER University
Udaipur, India
Priya Parsoya
Department of Chemistry Monika Trivedi
PAHER University Department of Chemistry
Udaipur, India PAHER University
Udaipur, India
Dmitry Polyansky
Ritu Vyas
Chemistry Department
Department of Chemistry
Brookhaven National Laboratory
Pacific Institute of Technology
Upton, New York
Udaipur, India

Pinki B. Punjabi Yasmin


Department of Chemistry Department of Chemistry
M. L. Sukhadia University Techno India NJR Institute of Technology
Udaipur, India Udaipur, India
1 Introduction
Suresh C. Ameta

CONTENTS
1.1 Energy Crisis and Solar Energy................................................................................................1
1.2 Solar Cells..................................................................................................................................3
1.3 Advantages.................................................................................................................................3
1.4 Disadvantages............................................................................................................................4
1.5 Future......................................................................................................................................... 4

1.1  ENERGY CRISIS AND SOLAR ENERGY


Solar radiation is the most abundant natural energy source for our planet. Solar radiation can
be transformed into various energy forms like heat and chemical energy. The sun provides
about 120,000 terawatts to Earth’s surface, which is 6000 times more than the present rate
of the world’s energy consumption; therefore, solar energy is going to play an important role
as a future energy source. Making use of this solar energy for generating environmentally
friendly chemical fuels and developing new and efficient photoconversion devices to convert
this energy into the form of electricity is a challenge for chemists, in general, and photochem-
ists, in particular. One of the advantages of a new generation of solar cells is their relatively
low cost.
The solar age began in the 1950s in Bell Laboratories with the development of silicon technol-
ogy. They described the first high-power silicon photovoltaic cell with 6% efficiency. This cell
was much more efficient than the previous selenium solar cell. Due to the energy crisis in 1970,
many technologies evolved for producing energy from renewable energy sources and photovoltaic
devices. A photovoltaic device or solar cell is a solid-state electrical device that directly converts
solar light energy into electricity. The light energy is transmitted in the form of small packets of
photons or quanta of light, and electrical energy generated is stored.
Energy is available as heat, hydropower, electricity, chemical, wind, geothermal, tidal, biomass,
liquid and gas fuels, sound, and many other forms, but there is a unity in the diversity of all of these
forms of energy. The potential energy of water in a dam is used to produce electricity in a hydroelec-
tric station, while chemical energy in fuels is converted into electrical energy or in a thermal power
station battery in a generator or an inverter.
Energy is a basic necessity for life. It has provided comfort to us in its different forms. It helps
us to cool down during summer and keeps us warm during winter. There has been a great increase
in the global demand for energy in the last few decades as a result of industrial revolution (devel-
opment) and population growth. It helps in transportation, construction, manufacturing, heat and
electricity production, mining, and so on.
There are three major challenges facing mankind in the coming few decades. These are water
shortages, energy challenges, and climate change. Of course, the latter depends on the first two. The
simultaneous existence of all of these problems seems nearly impossible to avoid. Energy remains a
great challenge to scientists, as the cost of solar energy is much higher than the cost of conventional
energy technologies.

1
2 Solar Energy Conversion and Storage

The world depends on fossil fuels at present, which are finite and not environmentally friendly.
Burning oil, petrol, diesel, and so on, releases harmful greenhouse gases and sometimes even car-
cinogens, resulting in pollution of the environment.
There have been a number of oil shocks in the last four decades. The first was in 1973–1974,
followed by those in 1979–1980, 1990–1991, 2002–2003, and the recent one in 2007–2008. It has
been assumed that this trend will continue in the future.
Since the dawn of the twenty-first century (also a new millennium), the generation of energy
from fuels, especially liquid fuels like diesel, petrol, and so on, with limitations on the rate of their
production, has created a bottleneck resulting in an energy crisis. An energy crisis is not actually
due to a shortage of energy sources, as there are enough energy sources available on Earth in some
form or other; it is due to our inability to extract sufficient fuel from the globe or to produce suf-
ficient electricity using available fuels.
The developed nations are facing such an energy crisis to meet the energy demands of the world.
This is caused by uncontrolled population growth. The burning of fossil fuel produces a large quan-
tity of the main culprit among greenhouse gases, carbon dioxide, which has an adverse effect on the
environment of the earth, causing global warming. There are two conflicting opinions about carbon
dioxide emissions. According to one, the world will be able to curb CO2 emissions to 490–535 ppm
by 2050, while the other is that it is able to reach 590–710 ppm. Both of these estimations are in
excess of the 450 ppm targets.
As a result, there is a threat of climate change, which can bring famines, droughts, diseases, and
so on. This is due to carbon emissions caused by inefficient, unclean uses of conventional energy
sources, which are nonrenewable.
Prices of oil and coal are soaring, and they have increased manyfold in the last few years. The
prices of other conventional energy resources are also skyrocketing. The increasing appetite for
energy developed in the last few decades has been further complicated by rapidly diminishing
conventional sources like coal and oil. This is added by the problem of ever-increasing demand and
constrained supply.
Fossil fuels like coal and oil have definitely played a critical role as a vast energy source, which
has fueled much growth in society, but now, it is agreed that these sources cannot continue to power
the development of society in the future. The rising demands for oil and decreasing supply are
likely to make fossil fuels economically unsustainable in years to come. Therefore, there is a press-
ing demand for renewable energy sources, which will decrease pollution in the environment. Solar
energy enters the scene here.
Solar energy is abundantly available, and it is practically inexhaustible. Biomass, wood, and food-
stuffs are derived from the sun. Even coal, petrol, diesel, and so on, are indirectly derived from the
sun. The sun radiates more energy per hour than the total energy consumed globally in 1 year. The
earth’s surface receives as much solar energy in just 1 year, which is two times the total reserves of
the earth’s conventional (nonrenewable) resources of oil, coal, natural gas, and uranium combined.
Solar energy is becoming increasingly popular as we face the problems caused by the burning of
conventional fuels such as wood, coal, kerosene, petrol, diesel, and so on. The popularity of solar
energy supported by increased efficiency and declining cost has grown to a level at which critics of
solar energy are becoming less vocal.
In an estimate made by the International Energy Agency (IEA), the energy needs of the world
will be 50% more in 2030 than at present. Another forecast is that worldwide energy demand could
increase almost threefold by the year 2050. Renewable energy has grown at a rapid pace in recent
years. According to an estimate, renewables produced 16.5% of primary energy requirements of the
world in 2005, which is a good sign as the proportion of renewable energy resources increases each
day and our dependence on nonrenewable energy resources decreases.
We face two main challenges: (1) generating electricity and (2) creating solutions for its stor-
age. Solar power accounts only for 34 TWh/y in the electricity generation mix, but this amount is
projected to increase by 225 times by 2050 (i.e., somewhere between 2980 and 7740 TWh/y). It has
Introduction 3

been estimated that solar energy currently contributes 4% of the world’s electricity. The prices of
fossil fuels are increasing at a rapid pace, while the prices of solar energy production have been
reduced by about 50% in the last decades. It is likely to become even cheaper in the years to come.
Common indicators of energy consumption are shares of electricity, heat, and transport fuels
from renewable energy resources. The share of renewable energy sources will continue to increase.
Some targets and scenarios indicate that the 20%–35% share of electricity from renewable by 2020
may increase to the range of 50%–80% by 2050.

1.2  SOLAR CELLS


Solar panels can convert sunlight into electrical energy (estimated 34% conversion efficiency,
theoretically); however, only about 24% has been achieved in the majority of panels so far.
The first generation of solar cell used in most devices utilizes silicon. Silicon is highly abundant,
and the best power conversion efficiency expected is 24%. But silicon solar cells are only 15%–20%
efficient. This means that 80%–85% of solar energy is lost as heat.
The III–V materials are more useful for photovoltaic applications as these are better absorbers of
light than silicon. These materials have the ability to stack multiple absorber layers, which provides
a better match between this absorber and the solar spectra; thus, the amount of loss of sunlight as
heat is reduced. As a result, the efficiency of the solar cell is increased.
The versatility and availability of different III–V materials make them quite attractive for pho-
tovoltaic applications. With these materials, one can harness different parts of the solar spectrum,
resulting in an overall increase in the efficiency of solar cells.
Nanostructured photovoltaic cells make use of electrical as well as optical properties of nano-
materials that can be controlled by changing the particle size at the atomic level, but the main chal-
lenge of incorporating these nanostructures into the solar cells without disrupting the fundamental
structure of the solar cell remains. There have been significant advances in the development of new
materials and devices that utilize nanotechnology and nanomaterials for the conversion and storage
of solar energy.
Quantum dots can be used to tune a solar cell and allow the absorption of the cell to be con-
trolled. The Nevada Policy Research Institute (NPRI) has shown enhanced efficiency in solar cells
with the use of indium (In), arsenic (As), gallium (Ga), and/or antimony (Sb) quantum dots in a stan-
dard gallium arsenide (GaAs) solar cell. They are able to show an increase in the current density.
Another approach is using concentrators in a photovoltaic cell by replacing III–V materials with
less costly materials. However, higher temperatures reduce the efficiency, but this added efficiency
from the concentrators outweighs the reduction in efficiency due to heat.

1.3 ADVANTAGES
• Solar energy is a clean and green source of energy, and it will never run out.
• It is estimated that all conventional energy reserves will be exhausted/depleted within a
few decades, while solar energy will not be exhausted for a long time.
• Solar energy does not depend on mining raw materials as is the case with petrol, diesel,
and natural gas, which causes forest and ecosystem restrictions.
• Solar energy not only benefits us financially but also protects (conserves) the environment.
• Solar energy systems are almost maintenance free, as a photovoltaic array will last for
decades once it is installed. Very little maintenance is required to keep solar cells function-
ing smoothly, as there are no moving parts in solar cells, while other conventional sources
used to generate electricity have moving parts.
• All conventional energy resources are costly, but solar energy is almost cost free except for
the costs of building and other equipment. Solar energy does not require any expensive raw
material, which is to be extracted, refined, and transported to the power plant.
4 Solar Energy Conversion and Storage

• It is quite expensive to transmit electrical energy, while solar energy can be used efficiently
even in remote areas.
• Conventional methods of generating energy cause pollution, whereas solar energy does not
cause any air pollution. However, the required equipment may cause some pollution during
manufacture. The major benefit of solar energy is that no greenhouse gases are generated;
therefore, it will not contribute to global warming, which results in climate change, sea
level rise, thinning of the ozone layer, and so on.
• Low power–consuming devices (calculators, mobile devices, watches, storage batteries,
etc.) may be charged effectively by solar energy devices.
• Solar energy is available in almost every part of the world (i.e., it is decentralized in many
locations), while conventional sources have deposits in some of the localized areas of the
world. Even in areas that are inaccessible for power transmission by power cables, solar
energy can produce electricity, no matter how remote the area, as long as sunlight is avail-
able there.
• Solar cells make absolutely no noise. There is not even a single beep when energy is har-
nessed, while the large machines used for extracting energy from coal, oil, petrol, and
diesel are extremely noisy.
• Solar energy is renowned for its versatility; therefore, it can be used to power any device,
from a small torch to a satellite.

As compared to a large number of advantages, solar energy has few disadvantages.

1.4 DISADVANTAGES
• Use of solar energy has its limits. It can be used only on sunny days. It cannot be properly
utilized in adverse conditions like storms, fog, nighttime, clouds, and so on. Solar energy
depends on regional, seasonal, and daily fluctuations, as solar energy is not constant in
various regions of the world or certain seasons of the year, and it also is not available at
night. In desert areas, large amounts of solar energy are available. In winter, the photo-
period of the day decreases as compared to summer days, and the intensity of the sun’s
radiation is also less.
• Large areas of land are required to capture solar energy. Areas of diffused sunlight require
a large number of solar panels that utilize more space and are quite expensive.
• Batteries are to be charged if solar energy is to be used at night. These batteries are larger in
size and heavy, so a space for storage is required, and these must be periodically replaced.
• At present, power generation from solar energy is quite expensive; however, the prices are
coming down with time.

1.5 FUTURE
Solar energy is a completely renewable energy resource. It will surely become affordable in the
future as low-cost materials are made available for producing electricity, but at present, electricity
produced from solar energy is quite expensive as compared to conventional methods. Recently, in
a world energy congress, it was stated that the world is going to face several significant challenges
in balancing global energy needs and addressing the energy “trilemma” (a balance between energy
security, social impact, and environmental sensitivity) over the next four decades.
In the future, efforts to develop new energy technology and increase the role of renewable and
innovative solutions of reducing pollution and greenhouse gases should be encouraged. Solar tech-
nology is improving and its cost is decreasing, while prices of fossil fuels are constantly increasing.
Improvements in the efficiency of conversion and storage of solar energy are required to reduce
dependence on variability and intermittence of solar power. There is also an existing challenge to
Introduction 5

reduce the cost of renewable energy per unit, and to meet this challenge, continuous effort should
be made to achieve enhanced efficiency of solar cells and bring down their costs. Energy transition
is not just an imperative, it is a certainty.
A complete replacement of energy presently derived from fossil fuels with energy from alterna-
tive renewable sources is almost impossible in the short term, and it may be unrealistic to expect this
transition even in the long term. If this transition is not properly managed, the consequences may be
quite severe. A meaningful energy transition is still more a theory than a reality.
The sun is a known, potentially huge source of clean and renewable energy. It has been estimated
that sunlight provides 10,000 times as much power as the world is using at the outset of the twenty-
first century. However, still there are some major technological challenges to be met in harnessing
this energy effectively. The efficiency and cost of harnessing solar power may not be ideal right now,
but with improvement, the future seems quite bright and sunny.
The energy crises we experience are the result of our own choices. Our energy future will be
decided by our own decisions and actions. There may be debate about the energy resources of the
future, but nobody can deny the fact that solar energy is the only future energy resource.
To achieve a bright future with sufficient energy to fulfill our present and future demands, let us
join the march into the solar age as soon as possible for a better world.
2 Photochemical Solar
Energy Conversion
Rakshit Ameta, Chetna Ameta, and Poonam Kumawat

CONTENTS
2.1 Solar Energy Scenario...............................................................................................................7
2.2 Some Photochemical Conversion Modes..................................................................................9
References......................................................................................................................................... 15

2.1  SOLAR ENERGY SCENARIO


The world is currently facing energy crises, and the problem will be more grave in the future as our
energy demands are ever increasing and the availability of energy sources and production of energy
are limited.
A decision has been made to increase the share of renewable energy sources to about 20% of
total production by 2020. No doubt, generation of energy in our present and future power plants
will have its own importance, but energy sources like coal, diesel, and so on, should be carefully
used and should not be wasted. These power sources are not very eco-friendly, but their significance
cannot be avoided until these resources are either completely or partially replaced by some alternate
renewable energy sources.
Sunlight is a clean source of energy that is inexhaustible and will remain low cost even in the
future; of course, presently, it is free of cost. In spite of several benefits, there are some limitations
in utilizing solar energy, such as night, fog, storms, and so on. There are also variations, such as
regional and seasonal variations, that limit the use of solar energy.
The source of energy in the Sun is thermonuclear reactions involving fusion of hydrogen to form
helium. About 4 billion years ago, the Earth was created, and the energy reaching the earth from
the Sun was responsible for the overall development and survival of life on this planet. Humankind
has used solar energy to fulfill their basic needs since ancient times.
A large amount of nuclear energy (3.89 × 1026 J s–1) is released by the core of the Sun as a result of
fusion reaction. This energy flux is converted in the form of thermal energy and transported toward
the surfaces of different planets. The total power density emitted by the Sun is of a very high order
(i.e., 64 MW m–2), but only a small part of it (i.e., of the order 1.366 kW [0.001366 MW m–2]) enters
just above Earth’s atmosphere. Therefore, with an assumption that there is no significant absorption
in space, this value, 13,766 W m–2, is termed the solar constant.
Solar radiation covers a large portion of the total spectral range. It includes radiation of nano-
metric and metric wavelengths. Overall solar energy flux is not equal in all ranges. It can be easily
divided into three major spectral categories:

• Ultraviolet (UV) radiation (λ < 400 nm) is less than 9%.


• Visible (VIS) radiation (λ between 400 and 800 nm) is around 39%.
• Infrared (IR) radiation (λ > 800 nm) is above 52%.

The solar spectrum is quite close to the radiation of a perfect black body at 5800 K. It indicates
air mass 0 (AM0) and air mass 1.5 (AM1.5), which are 1366.1 and 844 W m –2, respectively. The

7
8 Solar Energy Conversion and Storage

AM0 reference is for outside the terrestrial atmosphere; however, the radiation reaching there on
the surface may be affected by many factors. The major factors are lination of the Earth’s axis
and the atmosphere, which are responsible for both the absorption and repulsion (albedo) of this
incoming solar insolation. AM1.5 gives the influence of all these affecting parameters on solar
radiation. In this spectrum, the effect of molecular alignments present in the atmosphere on light
absorption is clearly visible.
The annual mean of solar radiation reaching the Earth’s surface is about 180 W m–2, and it is
slightly lower for the oceans (170 W m–2). This much solar radiation reaches the surface taking into
consideration the rotation of the Earth (day and night cycle); absorption by the contents of the atmo-
sphere; and reflection from cloud tops, oceans, and other terrestrial surfaces. Out of these, about
75% of light is received and roughly 25% is scattered by water vapor, air molecules, air Hodges,
clouds, and so on.
On a clear day, the solar insolation perpendicular to the surface of the Earth is approxi-
mately 1000 W m –2. This is the solar influx in clear weather at noon The solar influx is almost
the same all over the globe, in spite of the increased path through the atmosphere at higher
latitudes. It varies depending on the number of sun hours, clouds, and so on, in different parts
of the Earth.
Sunlight is diffused as well as refreshed. This value is important for the point of solar energy
conversion. Diffusion of sunlight in winter is about 35%, whereas it is only 15% in summer.
Some part of UV light from the Sun is absorbed by the ozone layer of the atmosphere. This
ozone layer acts as a protective covering for life on Earth, as UV light is harmful for many biologi-
cal processes. However, this layer is relatively thin over the equator. At some locations, an alarming
situation has been reached due to this thinning of the ozone layer, ultimately generating an ozone
hole due to various anthropogenic activities.
The Langley (L) is the conventional unit of the amount of solar radiation. One Langley is equal
to 1 calorie cm–2, and the incident power is normally expressed as Langleys per minute. The value
of a solar constant is 1.940 L min–1, which is 135.3 mW cm–2 or 423 Btu ft–2 h.
The intensity of the solar radiation reaching Earth’s surface depends on the distance traveled by
the light through the atmosphere. It is defined in terms of the air mass ratio (m) as
1
m= (2.1)
cosθ

where θ is the angle between the vertical direction and the direction of incident radiation.

• Air mass 0 (AM0) is the condition just outside the atmosphere.


• Air mass 1 (AM1) is the situation when the Sun is at the zenith (θ  =  90o  –  90o  =  0o;
cos 0° = 1 and m = 1/1 = 1).
• Air mass 1.5 (AM1.5) is used for radiation when the Sun is 41.8° to the horizon
(θ = 90° – 41.8° = 48.2°; cos 48.2° = 0.6666 and m = 1/0.6666 = 1.5).
• Air mass 2 (AM2) is used for radiation when the Sun is 30° to the horizon
(θ = 90° – 30° = 60°; cos 60° = 0.5 and m = 1/0.5 = 2).

The light intensity at a place is not much lower than this solar constant, if the Sun is almost overhead
atmospherically and it is a bright, sunny day.
Spectral distribution of sunlight can be represented in different ways. It may be a spectral pho-
tonic flux digital, N(λ), or a spectral irradiance, P (λ), against wavelength.
P(λ) and N(λ) are related as

⎛ hν ⎞
P ( λ ) = ⎜ ⎟ ⋅ N(λ) (2.2)
⎝λ⎠
Photochemical Solar Energy Conversion 9

This relation can also be expressed in terms of wave numbers, where P(ν) and N(ν) have the same
significance but in term of wave numbers,

P(hν) = (hcν)⋅ N(ν) (2.3)

The photonic flux density is the number of photons per unit area per unit time per unit wavelength
increment, while spectral irradiance is the energy per unit area per unit time per unit wavelength
increment. The total irradiation, Ptot, is presented as

∞ νm νm
Ptot = ∫ λm
P(λ)dλ = ∫ 0
P(ν) dν = ∫ 0
(hcν)⋅ N(ν)⋅ d(ν) (2.4)

The average value of the energy of a photon of white light 〈E〉 is

νm

E =
∫ 0
(hcν)⋅ N(ν)⋅ d(ν)
(2.5)
νm
Ptot
∫ 0
N(ν)⋅ d(ν) =
N tot

where Ntot is the number of photons (at all wave numbers) incident per unit area per unit time.

2.2  SOME PHOTOCHEMICAL CONVERSION MODES


The term solar fuel is normally used for a system involving any chemical process that captures
the light energy (photon energy), converts it, and stores it in the form of chemical energy (in the
form of a chemical bond). In the last five decades or so, efforts have been made to mimic natural
photosynthesis under laboratory conditions. Natural photosynthesis converts carbon dioxide of the
atmosphere into useful high-energy material like glucose. One thing that can be made clear here is
that in a glucose molecule, the carbon comes from carbon dioxide, and hydrogen and oxygen form
water molecules. It is a misnomer to say that plants convert carbon dioxide into oxygen. Actually,
this oxygen comes from water molecules, and it has been confirmed by isotopic labeling.

6CO 2 +12 H 2O* !hν


!!!" C H O +6O*2 +6H 2O
Chl. 6 12 6

In this process, all six carbon atoms of the carbon dioxide are incorporated in forming a glucose
molecule along with its six (out of 12) oxygen atoms. The 12 hydrogen atoms forming a water mol-
ecule are utilized in the formation of a glucose molecule, while the remaining 12 oxygen atoms
of water molecule and 6 oxygen atoms of carbon dioxide form 6 molecules of oxygen, which is
released.
This process of photosynthesis involves chlorophyll as a sensitizer. This photosynthetic reaction
looks quite simple, but it is mechanistically much more complex. Despite efforts by various chem-
ists and photochemists, artificially mimicking photosynthesis in the laboratory remains a challenge.
It is interesting to note that in natural photosynthesis, plants utilize only about 0.023% of solar
energy to feed the whole world. Looking at it this way, total solar insolation can be appreciated. The
solar fuels produced by solar energy are storable, as these are generated through biomass.
A lot of effort is going into mimicking photosynthesis under laboratory conditions. A question
arises in the minds of scientists or even laymen that when nature can convert carbon dioxide of the
10 Solar Energy Conversion and Storage

atmosphere into a useful component like biomass and produce oxygen, which is a very important
component for survival of life on Earth, why can an important and beneficial process not be carried
out artificially?
Chlorophyll is the natural sensitizer in photosynthesis, which can be either used as such or
replaced by some other sensitizer in artificial photosynthesis. However, there are still some prob-
lems in the conversion of carbon dioxide to solar fuels, like rates of reaction, cost effectiveness,
lifetime of various species involved, selectivity, overpotentials, and so on. Carbon dioxide (CO2) can
be reduced to useful synthetic fuels like HCOOH, HCHO, CH3OH, and ultimately CH4. However,
there are some reports of conversion of CO2 into these synthetic fuels but with very low efficiency.
It is a requirement that existing conventional energy sources be replaced by some alternate
renewable, abundant, clean, and readily available energy source. This ambition can be fulfilled by
artificial photosynthesis, which will utilize solar energy, CO2, and water to generate biomass, but
much more effort is required to achieve this goal.
A dye-sensitized solar cell (DSSC) is a type of thin-film solar cell. It is also known as the Gratzel
cell in honor of Michal Gratzel, who invented it with Brian O’Regan. Later, it was further developed
at Lausanne, Switzerland (O’Regan and Gratzel 1991). It provides a technically and economically
viable concept as an alternative to p-n junction photovoltaic devices. It is alternatively abbreviated
as DSSC, DSC, or DYSC. It has a number of attractive features, for instance, it is semiflexible and
semitransparent, offering a variety of applications not possible with a glass-based system. In addi-
tion, most of the materials used in DSSCs are relatively low cost.
Normally, a semiconductor is responsible for both light absorption and charge carrier transport
in conventional systems. These two processes are separated in a dye-sensitized solar cell. A sen-
sitizer absorbs the light, which is anchored to the surface of a wideband semiconductor, whereas
charge separation is at the interface from the dye molecule into the conduction band of the solar cell
through a photoinduced electron injection. These charge carriers are then transported in the con-
duction band of the semiconductor, the charge collector. As the sensitizers have a broad absorption
band, their use in conjunction with oxide films of a nanocrystalline nature will permit us to utilize
a major portion of sunlight.
Dye-sensitized solar cells generate photocurrents from electron injection by the sensitizer dye.
These cells have a photoanode and are called n-DSSCs. The reverse is the case with a photocathode
(p-DSSC), which operates in the opposite manner to conventional DSSCs. Here, dye excitation is
followed by a rapid electron transfer from a p-type semiconductor to the dye. It is dye-sensitized
hole injection, while in n-type DSSC, it is electron injection. Some tandem solar cells (p-n-DSSCs)
can be constructed using a combination of such n-DSSC and p-DSSC. The calculated/theoretical
efficiency of these tandem DSSCs is far beyond that of a single-junction n-type or p-type DSSC.
Any tandem cell (p-n-DSSCs) will have one n-DSSC and one p-DSSC in a simple sandwich form
with an intermediate electrolyte layer. These n-DSSCs and p-DSSCs are connected in series. This
implies that a resulting photocurrent will be controlled by the weakest photoelectron, and the pho-
tovoltages are additive. Therefore, it is very important to have photocurrent matching for the con-
struction of any tandem DSSCs with high efficiency; however, fast charge recombination following
dye-sensitized hole injection (unlike n-DSSC) will result in relatively low photocurrent in p-DSSCs.
Thus, the efficiency of the overall device is reduced.
The first DSSC was designed in three basic parts. The first part is a transparent anode made of
fluorine-doped tin oxide (FTO) deposited on the back of a glass plate. This conductive plate has a
thin layer of titanium dioxide on the back, which is in the form of a highly porous structure with high
surface area. It is known that titanium dioxide absorbs only a small fraction of solar insolation and
that, too, in the UV range. This plate is then immersed in a mixture of a photosensitive ruthenium–
polypyridine complex (also called molecular sensitizer) and a solvent. A thin layer of the dye was left
covalently bonded on the surface of titanium dioxide after soaking the film in the dye solution. Then
a separate plate was made with a thin layer of iodide electrolyte spread over a conductive sheet of
the metal platinum. These two plates are joined and sealed together so that there is no leakage of the
Photochemical Solar Energy Conversion 11

electrolyte. The efficiency of the cell depends on four energy levels: (1) the LUMO (conduction band),
(2) the HOMO (valence band) of the photosensitizer, (3) the Fermi level of the TiO2, and (4) the redox
potential of the mediator (I−/I3−) in electrolyte. Sunlight enters the DSSC through transparent FTO
top contact and strikes the dye molecule on the surface of the TiO2. Those photons that have enough
energy to be absorbed excite the dye molecule to its excited state. An electron from the excited state
of the dye is injected directly into the conduction band of TiO2. This electron diffuses to the anode
on top as a result of an electron concentration gradient. If another electron is not available, then the
dye molecule losing an electron will decompose.
The dye molecule will abstract an electron from iodide in the electrolyte below TiO2. This will
lead to oxidation of iodide into triiodide (I3−). This electron transfer occurs quite rapidly compared
to the time taken for the injected electron to recombine with the oxidized dye molecule. Thus, a
recombination reaction is prevented, and the solar cell is short circuited. The triiodide then diffuses
mechanically to the bottom of the cell, recovering its missing electron, where the counterelectrode
will reintroduce the electrons after flowing to the external circuit.
Organic dyes have been used as sensitizers along with phthalocyanines, porphyrins, and
so on, and high solar-to-electric power conversion efficiencies have been achieved using these
sensitizers. Semiconductor quantum dots are also good sensitizers. These quantum dots are
mostly II–VI and III–V types of semiconductor particles, and their absorption spectra can be
adjusted by variation of their particle sizes. However, one of the problems with quantum dots is
photocorrosion.
One of the main components of DSSC is electrolytes, as it can affect stability as well as the con-
version efficiency of the solar cell.
There are three types of electrolytes used in DSSCs:

• Liquid electrolyte
• Solid electrolyte
• Quasi-solid-state electrolyte

Liquid electrolytes can be further classified into organic solvent electrolytes and ionic liquid electro-
lytes. Organic solvent can be used as liquid electrolyte (e.g., acetonitrile, ethylene carbonate, valero-
nitrile, etc.). The ionic liquid electrolyte can also be used. 1-Methyl-3-alkyl-imidazolium iodide or
1-methyl-3-ethyl/propyl/butyl imidazolium salts have been most commonly used as ionic liquid
electrolytes. Various counterions have been used in these salts. These are I−, BF4−, NCS−, PF6 −,
and so on. Ionic liquids are commonly used in DSSCs because of their good stability, high ther-
mal stability, high ionic conductivity, negligible vapor pressure, and so on. Various redox couples
have been used as electrolytes. These are Br−/Br2, I3−/I−, SCN−/(SCN)2, SeCN−/(SeCN)2, and so on
(Boschloo and Hagfeldt 2009; Wang et al. 2010).
Some solid-state electrolytes have been used from time to time in DSSC, such as 1-methyl-
3-acetyl-imidaolium iodide (Zhao et al. 2008), 4-cyano-4′-hydroxybiphenyl and imidazolium units
(Cao-Cen et al., 2012), [((3-(4-vinyllpyridine) propanesulfonic acid) iodide)-co-(acrylonitrile)]
(Fang et al. 2011), and carbazole-imidazolium cation (Midya et al. 2010), with efficiency varying
from 2.85% to 6.95%.
Similarly, some quasi-solid-state electrolytes were also used in dye-sensitized solar cells. Kumara
et al. (2002) used triethylamine hydrothiocyanate as a CuI crystal growth inhibitor, while room-
temperature molten salt, 1-hexyl-3-methylimidazolium iodide, iodine, and low-molecular-weight
gelator were used by Kubo et al. (2002) with 5% conversion efficiency. An elastomeric copolymer of
ethylene glycol and epichlorohydrin was employed in DSSC (Nogueira et al. 2001). Good efficiency
of 5.4% was observed by Stathatos et al. (2003) using sol-gel nanocomposite electrolyte containing
a surfactant Triton X-100, propylene carbonate, and 1-methyl-3-propylimidazolium iodide.
Although quasi-solid-state ionic liquid electrolytes have a better stability, their conversion effi-
ciency is not as good as that of liquid electrolyte.
12 Solar Energy Conversion and Storage

The concept and fundamental operation of a photogalvanic cell is quite different from that of
a photovoltaic cell. These are based on chemical reactions that give rise to energy-rich products
on excitation by a photon. Thereafter, these high-energy products lose energy electrochemically.
Thus, photogalvanic processes are defined as different types of physical and chemical processes
converting a flux of light energy into electrical power. A photogalvanic device functions as a simple
transducer or it may store sufficient energy as chemical potential under an open-circuit condition
and release this energy as electricity on closing the external circuit. The first photogalvanic cell was
an iron–thionine cell, where a platinum electrode was illuminated and the other counterelectrode
was kept in the dark.
This system has five basic steps:

1. Absorption of incident light by a colored material (dye/complex)


2. Conversion of this energy into chemical potential of a charge carrier
3. Diffusion of these charge carriers to the surface of the electrode
4. Transfer of the charge to the surface of the electrode
5. Generation of current in an external circuit

Various modifications to the photogalvanic cell have been made in the last few decades, such as
modification of the electrode, synthesis, and the use of newer colored systems, exposing alterna-
tively the two electrodes, use of micelles, and so on.
The photoelectric effect was discovered by Becquerel. Since that discovery, efforts have been
made to convert sunlight to electric energy. When a photon strikes a semiconductor, it can create an
electron–hole pair, and as a result, an electric potential develops.
To simplify, the mechanism of a photoelectrochemical cell is based on this conversion involving
two electrodes. There are basically three possible photoelectrodes in the fabrication of a photoelec-
trochemical cell:

• A photoanode (n-type semiconductor) and cathode (metal)


• A photoanode (n-type semiconductor) and a photocathode (p-type semiconductor)
• A photoanode (p-type semiconductor) and a cathode (metal)

Initially, titanium dioxide and some other metal oxides were used in a photoelectrochemical cell
to achieve reasonable efficiency. Later, some titanates were used, where the conduction band was
mainly 3d character and the valence band was oxygen 2p character.
The bandgap of titania (approximately 3 eV) is wide enough, making it unable to absorb visible
radiation more efficiently. Various modifications have been made to bring the absorption in the vis-
ible range to make it more effective. TiO2 nanowires over porous nanocrystalline TiO2 have also
been explored for their possible use in a photoelectrochemical cell. Gallium nitride as well as other
metal nitrides can also be good options because they can utilize almost the entire solar spectrum as
their bandgaps are narrow enough (Wang et al. 2011). Some other nonoxides semiconductors have
also been used, such as MoS2, MoSe2, WSe2, GaAs, and so on, as n-type electrodes (Kline et al.
1981).
The field of photoelectrochemistry has been discussed by different people (Fujishima et al. 1969;
Cao et al. 1996; Tyrk et al. 2000). One of the major applications of a photoelectrochemical cell is
water cleavage. It provides generation of synthetic fuel, hydrogen. Hydrogen has been advocated as
the fuel of the future.
The synthesis of energy-rich molecules is one of the main objectives in the chemical storage of
solar energy. These energy-rich molecules may be formaldehyde, formic acid, methanol, methane,
hydrazine, and so on, so that they can store energy, and their long-range distribution is also possible.
The reduction of carbon dioxide could be a good alternative as it could solve both the prob-
lems: the generation of synthetic fuels to solve the problem of energy crises and also to control the
Photochemical Solar Energy Conversion 13

ever-increasing amount of carbon dioxide in the atmosphere. Conventional energy fuels (i.e., diesel,
petrol, kerosene, coal, etc.) add carbon dioxide to the atmosphere, resulting in global warming and
ultimately natural catastrophes, while the use of synthetic fuels prepared by the reduction of carbon
dioxide would not add even a single molecule of carbon dioxide to the atmosphere as these fuels are
generated by photoreduction of carbon dioxide. Second, formaldehyde is the photoreduced product
in some cases. It can be used as a precursor for the synthesis of glucose, which is a step toward
mimicking photosynthesis.
Hydrogen is the most useful form of converted solar energy, as it can be easily substituted for
petroleum-based fuels. Photodissociation of water in natural photosynthesis is an endothermic reac-
tion, and 284 kJ mol−1 energy is released on decomposition of water in the form of hydrogen and
oxygen. For this conversion, 1.23 eV of energy are required.
The system for generation of H2 from the photosplitting of water consists of three components:

• Compounds include two components, that is, reductant and oxidant, which can be reduced
or oxidized by quenching the excited species.
• A catalyst, which is able to collect the electrons and transfer them to the water.

However, a fourth species, electron donor, may also be required to prevent back-electron transfer or
recombination of an electron–hole pair.
Hydrogen has its own importance as a synthetic fuel because of its large storage capacity, (i.e.,
approximately 119,000 J g−1). Hydrogen can generate electricity, a source of energy, when burned in
a fuel cell, and produces water as the product. Water, on photolytic decomposition, gives hydrogen
and oxygen utilizing only light energy. These two chemical components, hydrogen and oxygen, are
original sources and will give energy without polluting the environment. So photogeneration of
hydrogen is welcome, as it can fulfill our energy demands of the future because it is renewable in
nature, while other conventional fuels not only harm the environment but are limited and cannot
be regenerated.
Photogeneration of another molecule is also interesting from the solar energy storage point
of view. Of interest is photosynthesis of hydrogen peroxide from an oxygen molecule. It was
reported that ethanol in the presence of oxygen and alkaline medium produces hydrogen perox-
ide in the presence of metal-free phthalocyanine or eosin. Similarly, regeneration of free halogen
from halide ions has its own importance in this context. One interesting reaction is the formation
of chlorine from Cl− in the presence of sulfonated anthraquinone salts. The oxidation of chloride
to chlorine is also endothermic in nature. The formation of a strong oxidant may be considered
as a novel way to store solar energy. Another promising system of storing solar energy is pho-
toisomerization of norbornadiene to quadricyclane (although it has low enthalpy), dimerization
of anthracene, ring-opening reaction of piperidine and pyrrolidine derivatives, and so on. The
reduction of nitrogen to its reduced counterparts, like hydrazine, may prove to be a reaction of
interest from an energy point of view. Photolysis of nitrosyl chloride and the decomposition of
nitrogen dioxide, phosgene, and sulfur dioxide are other reactions of importance.
A photovoltaic cell is a device that converts visible light into direct current. Nanomaterial
absorbs a photon, and it creates an excited state. This excited state is an electron–hole pair bound
together by the electrostatic interaction. These are called excitons. Excitons are then broken up into
free electron–hole pairs by some effective field in photovoltaic cells. An effective field is set up by
creating a heterojunction between two different (dissimilar) materials. This effective field breaks
up excitons as electrons fall from the conduction band of the absorber to the conduction band of the
acceptor molecule. For this transfer, the conduction band of the absorber material needs to be higher
than the conduction band edge of the acceptor material (McGehee and Topinka 2006; Nelson 2002;
Halls and Friend 2001).
The simplest example of an organic photovoltaic cell is a monolayer or single-layer type. In
these cells, a layer of organic electronic material is sandwiched between two metallic conductors.
14 Solar Energy Conversion and Storage

Normally, a layer of indium tin oxide (ITO) is used, which has high work function. Another layer is
of metals like magnesium, aluminum calcium, and so on, which have low work function. Because
of the difference in these work functions, an electric field is set up in the organic layer. When this
organic layer absorbs light photons, electrons from the highest occupied molecular orbital (HOMO)
will be excited to the lowest unoccupied molecular orbital (LUMO), thus leaving behind a hole. The
potential created by these two different work functions helps to split them to exciton pairs. As a
consequence, electrons are pulled to the positive electrode and the holes to the negative electrodes.
Kearns and Calvin (1958) reported the photovoltaic effect with a photovoltage of 200 mV. A macro-
cyclic compound magnesium phthalocyanine was used by them. Ghosh et al. (1974) reported an Al/
phthalocyanine/Ag Schottky barrier cell. The cell showed photovoltaic cell efficiency of 0.01% on
illuminating it with radiation of 690 nm.
Weinberger et al. (1982) utilized polyacetylene as the organic layer sandwiched between alumi-
num and graphite. They produced open-circuit voltage of 300 mV with a charge collection efficiency
of 0.3%. The low quantum efficiency as well as low power conversion efficiency limit the use of
single-layer organic solar cells. Another problem is that the resulting electric field between two con-
ductive electrodes is not sufficient to split the excitons as electrons often recombine with the holes
before these reach the counterelectrode.
Another type of organic photovoltaic cell is bilayer cells, with the only difference being that
there are two layers sandwiched between two electrodes. One of the layers is an electron-donating
layer, and the other is an electron-accepting layer. As these two layers have different electron affin-
ity and ionization energy, electrostatic forces are generated, and the interface between these two
layers is relatively large. The local electric fields are therefore stronger, which splits excitons more
efficiently than single-layer photovoltaic cells. Such a structure is also known as a planar donor–
acceptor hydrogen. A poly(p-phenylene vinylene) (PPV/C60) photovoltaic cell was fabricated by
Halls et al. (1996). They reported quantum efficiency of 9%, power conversion efficiency of 1%, and
a fill factor of 0.48%. They also reported 6% quantum efficiency and a fill factor of 0.6 using a cell
with a layer of bis(phenethylimido)perylene over a layer of PPV (Halls and Friend, 1997). In such a
cell, the problem is diffusion of excitons to the interface of the layers and their splitting into carriers.
Approximately 100 nm thickness of polymer layer is required to absorb enough light. Thus, only a
limited fraction of excitons are able to reach the heterojunction interface.
Imec (Belgium) fabricated a photovoltaic cell with 8.4% efficiency using a three-layer (two elec-
tron acceptors and one donor) fullerene free stack. The open-circuit voltage of 1 V was achieved.
Different kinds of heterojunctions have been used in photovoltaic cells, such as discrete, bulk, and
gradient heterojunctions.
Some other types of solar cells are fabricated either by modification of existing solar cells or by
a combination of some solar cells. These are hybrid solar cells, biohybrid solar cells, plastic solar
cells, plasmonic solar cells, perovskite solar cells, and so on.
Biohybrid solar cells consist of a combination of inorganic matter and organic matter (photo-
system I). This type of solar cell has been prepared by research workers at Vanderbilt University,
Nashville, Tennessee. Ciesielski et al. (2010) prepared photosystem I biohybrid photoelectro-
chemical cells. In these cells, photosystem I complexes on the surface of a cathode will generate
photocurrent density (approximately 2  μA  cm–2) through a photocatalytic effect. These biohy-
brid solar cells have remarkable stability and remain active for more than 9 months. Yehezkeli
et al. (2012) used integrated photosystem II–based photobioelectrochemical cells, which generate
electricity on illuminating biomaterial functionalized electrodes in aqueous solutions. The pho-
toanode used was photosystem II functionalized and electrically wired bilirubin oxidase/carbon
nanotubes modified as cathode. Here, water is oxidized to O2 at the anode, while O2 is reduced
to water at the cathode.
Similarly, hybrid solar cells utilize organic as well as inorganic semiconductors. In this case, the
organic material, which is normally conjugated polymer, absorbs light as the donor and as a con-
sequence, before the hole is transferred. The inorganic materials are used in these cells as electron
Photochemical Solar Energy Conversion 15

transporters. Various electron acceptors used in such devices are fullerenes, polymers, inorganic
nanocrystals, and so on. Thick oxide nanoparticles dispersed in a semiconducting polymer were
used by Beek et al. (2004), which act as an active layer converting 40% of incident light into electri-
cal power at 500 nm with conversion efficiency of about 1.5%. It may be considered a step toward
green electricity if eco-friendly materials are used at low temperatures. Beek et al. (2005) also
reported such hybrid solar cells using crystalline zinc oxide nanoparticles as the electron acceptor
and conjugated polymer blend of poly [2-methyl-5-(3′,7′-dimethyloctyloxy)-1,4-phenylene vinylene]
(MDMO-PPV) as the electron donor. They also investigated the effect of degree of type of mixing
of the two components as well as size and shape of nanocrystalline zinc oxide particle and observed
optimized performance of about 1.6% and photon-to-current conversion efficiency ~50%.
The light energy may also be converted into electricity using plasmons. Such solar cells are
called plasmonic cells. These cells are thin-film solar cells, where thickness of the film is about
1–2 μm. Here, lower-cost materials than silicon can be used, such as plastic, glass, steel, and so on.
Catchpole and Polman (2008) have reviewed plasmonic solar cells. They reported that scattering
from metal nanoparticles is a promising way of increasing light absorption in thin solar cells, which
is commonly encountered in thin solar cells. Brown et al. (2010) reported the synthesis of core–shell
metal insulator nanoparticles using Au-SiO2-TiO2 and plasmonic enhancement of dye-sensitized
solar cells. These particles provide better efficiency than Au-SiO2 plasmonic nanoparticles. It was
also concluded that this is due to a near-field plasmonic effect (Sheehan et al. 2013).
The use of polymer-based photovoltaic materials makes it possible to obtain low-cost materi-
als. The photoinduced electron transfer from donor-type semiconducting material with acceptor-
type polymers, fullerenes, and so on, is used in these plastic solar cells. However, Sariciftci (2004)
reported that power conversion efficiency ~5% is possible in such plastic photovoltaic devices. It was
anticipated that this will increase by 8%–10% in the future.
Some inorganic and hybrid light absorbers can also be used in thin-film photovoltaic devices like
quantum dots and organometal halide perovskite as these have good potential for high conversion
efficiency and are also low in cost. It was demonstrated that by using meso-superstructure organo-
metal halide perovskite solar cells with core shell Au-SiO2 nanoparticles, the efficiency of the solar
cell could be improved to 11.4%. This opens up an avenue for facilitated tuning of exciton binding
energies in perovskite semiconductors. Chiechi et al. (2013) have reviewed the field of plastic solar
cells and also the trends in polymeric nanomaterials and fullerene acceptors for hybrid polymer/
quantum dot devices.

REFERENCES
Beek, W. J. E., M. M. Wienk, and R. A. J. Janssen. 2004. Efficient hybrid solar calls from zinc oxide nanopar-
ticles and a conjugated polymer. Adv. Mater. 16: 1009–1013.
Beek, W. J. E., M. M. Wienk, M. Kemerink, X. Yang, and R. A. J. Janssen. 2005. Hybrid zinc oxide conjugated
polymer bulk heterojunction solar cell. J. Phys. Chem. B. 109: 9505–9516.
Boschloo, G., and A. Hagfeldt. 2009. Characteristics of the iodide/triiodide redox mediator in dye-sensitized
solar cells. Acc. Chem. Res. 42: 1819–1826.
Brown, M. D., T. Suteewong, R. S. S. Kumar, V. D’Innocenzo, A. Petrozza, M. Lee, et al. 2010. Plasmonic
dye-sensitized solar cells using core-shell metal-insulator nanoparticles. Nano Lett. 11: 438–445.
Cao, F., G. Oskam, G. J. Meyer, and P. C. Searson. 1996. Electron transport in porous nanocrystalline TiO2
photoelectrochemical cells. J. Phys. Chem. 100: 17021–17027.
Cao-Cen, H., J. Zhao, L. Qiu, D. Xu, Q. Li, X. Chen, et al. 2012. High performance all-solid-state dye sen-
sitized solar cells based on cyanobiphenyl-functionalized imidazolium type ionic crystals. J. Mater.
Chem. 22: 12842–12850.
Catchpole, K. R., and A. Polman. 2008. Plasmic solar cells. Opt. Express. 16: 21793–21800.
Chiechi, R. C., R. W. A. Havenith, J. C. Hummelen, L. J. A. Koster, and M. A. Loi. 2013. Modern plastic solar
cells: Materials, mechanisms and modeling. Mater. Today. 16: 281–288.
Ciesielski, P. N., F. M. Hijazi, A. M. Scott, C. J. Beard, K. Emmett, S. J. Rosenthal, et al. 2010. Photosystem
1-baded biohybrid photoelectrochemical cells. Biosource Technol. 101: 3047–3053.
16 Solar Energy Conversion and Storage

Fang, Y., W. Xiang, X. Zhou, Y. Lin, and S. Fang. 2011. High performance novel acidic ionic liquid polymer/
ionic liquid composite polymer electrolyte for dye-sensitized solar cells. Electrochem. Commun. 13:
60–63.
Fujishima, A., K. Honda, S. Kikuchi, and K. K. Zasshi. 1969. Recent topics in photoelectrochemistry.
Achievements and future prospects. Electrochim. Acta. 72: 108–113.
Ghosh, A. K., D. L. Morel, T. Feng, R. F. Shaw, and C. A. Rowe Jr. 1974. Photovoltaic and rectification proper-
ties of Al∕Mg phthalocyanine∕Ag Schottky-barrier cells. J. Appl. Phys. 45: 230–236.
Halls, J. J. M., and R. H. Friend. 1997. The photovoltaic effect in a poly(p-phenylenevinylene)/perylene hetero-
junction. Synth. Met. 85: 1307–1308.
Halls, J. J. M., and R. H. Friend. 2001. In Archer M. D., and R. D. Hill (Eds.), Clean Electricity from
Photovoltaics. London: Imperial College Press, pp. 377–445.
Halls, J. J. M., K. Pichler, R. H. Friend, S. C. Moratti, and A. B. Holmes. 1996. Exciton diffusion and dis-
sociation in a poly(p-phenylenevinylene)/C60 heterojunction photovoltaic cell. Appl. Phys. Lett. 68:
3120–3122.
Kearns, D., and M. Calvin. 1958. Photovoltaic effect and photoconductivity in laminated organic systems.
J. Chem. Phys. 29: 950–951.
Kline, G., K. Kam, D. Canfield, and B. Parkinson. 1981. Efficient and stable photoelectrochemical cells con-
structed with WSe2 and MoSe2 photoanodes. Sol. Energy Mater. 4: 301–308.
Kubo, W., T. Kiramura, K. Hanabusa, Y. Wada, and S. Yanagida. 2002. Quasi-solid-state dye-sensitized solar
cells using room temperature molten salts and a low molecular weight gelator. Chem. Commun. 374–375.
Kumara, G. R., S. Kaneko, M. Okuya, and K. Tennkone. 2002. Fabrication of dye-sensitized solar cells using
triethylamine hydrothiocyanate as a Cul crystal growth inhibitor. 18: 10493–10495.
McGehee, D. G., and M. A. Topinka. 2006. Solar cells: Pictures from the blended zone. Nat. Mater. 5: 675–676.
Midya, A., Z. Xie, J. X. Yang, Z. K. Chen, D. J. Blackwood, J. Wang, S. Adams, and K. P. Loh. 2010. A new
class of solid state ionic conductors for application in all solid state dye sensitized solar cells. Chem.
Commun. 46: 2091–2093.
Nelson, J. 2002. Organic photovoltaic films. Curr. Opin. Solid State Mater. Sci. 6: 87–95.
Nogueira, A. F., J. R. Durrant, and M. A. De Paoli. 2001. Dye-sensitized nanocrystalline solar cells employing
a polymer electrolyte. Adv. Mater. 13: 826–830.
O’Regan, B., and M. Grätzel. 1991. A low-cost, high-efficiency solar cell based on dye-sensitized colloidal
TiO2 films. Nature 353: 737–740.
Sariciftci, N. S. 2004. Plastic photovoltaic devices. Mater. Today 7: 36–40.
Sheehan, S. W., H. Noh, G. W. Brudvig, H. Cao, and C. A. Schmuttenmaer. 2013. Plasmonic enhancement of
dye-sensitized solar cells using core-shell-shell nanostructures. J. Phys. Chem. C. 117: 927–934.
Stathatos, E., P. Lianos, S. M. Zakeeruddin, P. Kiska, and M. Gratzel. 2003. A quasi-solid-state dye-sensitized
solar cell based on a sol-gel nanocomposite electrolyte containing ionic liquid. Chem. Mater. 395:
583–585.
Tryk, D., A. Fujishima, and K. Honda. 2000. Recent topics in photoelectrochemistry achievements and future
prospects. Electrochim. Acta. 45: 2363–2376.
Wang, D., A. Pierre, M. G. Kibria, K. Cui, X. Han, K. H. Guo, et al. 2011. Wafer-level photocatalytic water
splitting on GaN nanowire arrays grown by molecular beam epitaxy. Nano Lett. 11: 2353–2357.
Wang, M., N. Chamberland, L. Breau, J.-E. Moser, R. Humphry-Baker, B. Marsan, et al. 2010. An organic
redox electrolyte to rival triiodide/iodide in dye-sensitized solar cells. Nat. Chem. 2: 385–389.
Weinberger, B. R., M. Akhtar, and S. C. Gau. 1982. Polyacetylene photovoltaic devices. Synth. Met. 4: 187–197.
Yehezkeli, O., R. Tel-Vered, J. Wasserman, A. Trifonov, D. Michaeli, R. Nechushtai, et al. 2012. Integrated
photosystem II-based photoelectrochemical cells. Nat. Commun. 3: Art. No. 742.
Zhao, Y., J. Zhai, J. He, X. Chen, L. Chen, L. Zhang, et al. 2008. High performance all-solid-state dye sen-
sitized solar cells utilizing imidazolium-type ionic crystal as charge transfer layer. Chem. Mater. 20:
6022–6028.
3 Basic Photoelectrochemistry
Purnima Dashora, Meenakshi Joshi, and Suresh C. Ameta

CONTENTS
3.1 Electrochemistry of Semiconductors....................................................................................... 17
3.2 Photoelectrochemistry.............................................................................................................24
3.3 Photocatalysis..........................................................................................................................26
References.........................................................................................................................................28

3.1  ELECTROCHEMISTRY OF SEMICONDUCTORS


Materials are normally classified into three categories, depending on energy of conduction band
(CB) and valence band (VB) (Figure 3.1). The conduction band is the lowest unoccupied molecular
orbital (LUMO), while the valence band is the highest occupied molecular orbital (HOMO). The
conducting materials have a bandgap less than 1.0  eV or, sometimes, there may even be almost
an overlap between these two bands, so that there is smooth conduction of electricity. The insula-
tors (nonconducting material) have a wide bandgap that is more than 5.0 eV; therefore, they do
not conduct electricity. There are some materials that have a bandgap of the medium order that is
1.5–3.0 eV. These are basically nonconducting in nature, but in the presence of some energy source,
they become conducting. Therefore, such materials are called semiconductors.
A metal (M) can be easily described as the number of metal ions (Mn+) arranged rigidly in a crys-
tal lattice and freely moving electrons. The number of electrons will be n times the number of metal
ions. These electrons are conducting in nature and distributed in different energy levels according
to the laws of quantum machines and statistical thermodynamics. At absolute zero temperature
(T = –273.15°C or 0 K), all the energy levels are filled up to a level called the Fermi level (Ef ). This
Fermi level is a characteristic of a particular material. There is a probability of finding electrons
in energy levels higher than Fermi levels and an equal number of holes in energy levels lower than
Fermi levels (Figure 3.2).
In the case of an insulator, all electrons are firmly attached to the atomic nuclei in the lattice or to
the region between these (valence electrons). As there are no free moving electrons in an insulator,
they show low or no conductance. However, if by any means it is possible to promote an electron
from the VB to CB resulting in the creation of a hole in the VB, then both of these carriers, electrons
and holes, are mobile under the application of the electric field. As a result, there is a possibility of
a slight increase in the electrical conductivity of that insulator.
A semiconductor in a pure form (with no impurities) is called an intrinsic semiconductor. The
number of electrons in the CB (per unit volume) equals the number of holes in the VB (per unit
volume). If the temperature and bandgap are known, then the values of the numbers of electrons and
holes can be calculated using Fermi–Dirac statistics.
The Fermi level in an intrinsic semiconductor is exactly half the way between the top of VB and
the bottom of CB (Figure 3.3).
Doping an intrinsic semiconductor by some impurities with a small amount will change the
property of that intrinsic semiconductor. The addition of an impurity of atoms with more valence
electrons than the atom of an intrinsic semiconductor will add to the number of electrons in doped
semiconductor than intrinsic semiconductor. Such doped semiconductors are called n-type or
n-doped semiconductors (Figure 3.4). In such cases, electrons are more than the holes. Therefore,

17
18 Solar Energy Conversion and Storage

CB
CB
CB

Energy
Eg < 1.0 eV Eg ≃ 1.5–3.0 eV Eg > 5.0 eV

VB
VB
Metal VB
Semiconductor
Insulator

FIGURE 3.1  Classification of materials.

T>0

EF T=0


FIGURE 3.2  Distribution of electrons and holes in energy levels at
NE


different temperatures.

E
Conduction
electrons
CB

T>0
Bandgap

Fermi level

VB Holes

FIGURE 3.3 Fermi level, electrons and holes in intrinsic


NE

semiconductor.

electrons will act as majority carriers, and holes will act as minority carriers. Here, the Fermi level
is close to the bottom of the CB, because the number of electrons is higher.
The reverse is true in the case of doping by an atom that has a number of electrons less than the
number of atoms of an intrinsic semiconductor. If this is the case, then the doped semiconductor
will be named p-type or p-doped semiconductor (Figure 3.5). Here, holes are majority carriers,
while electrons are minority carriers because the number of holes (nh) is greater than the number of
electrons (ne). As the number of holes is higher, the Fermi level is closer to the top of the VB.
At room temperature, the intrinsic semiconductor silicon has a number of electrons or holes
equal to approximately 3.0 × 1016 m−3, which becomes approximately 1023 m–3(n) and 1010 m–3(p) for
n-doped Si; here, n ≫ p. On the other hand, if silicon is doped with gallium (which has three valence
electrons only), then it gives p-doped Si, where p ≫ n. However, for the given semiconductor at a
certain temperature, the product of n and p is independent of degree of doping.
If a semiconductor is illuminated with an energy, which is more than its bandgap, then it was
observed that there is an increase in the conductivity of that semiconductor. This phenomenon is
known as photoconductivity. In the case of an intrinsic semiconductor, the number of electrons
Basic Photoelectrochemistry 19

E
Conduction
electrons

CB

EF Fermi level

Holes

VB

nE

FIGURE 3.4  n-Doped semiconductor.

E
Conduction
electrons
CB

EF

Fermi level Holes


VB

nE

FIGURE 3.5  p-Doped semiconductor.

and holes both increase equally. This is not the case with n-type and p-type semiconductors. In
such cases, there is a large increase in the number of minority carriers and a relatively insignificant
increase in the number of majority carriers. The photoexcitation of the n-doped semiconductor will
cause a large increase in the number of holes and a small increase in the number of electrons. The
reverse is true for the p-doped semiconductor, where there is a greater increase in the number of
electrons and a small increase in the number of holes.
If a metal comes in contact with an n-type semiconductor, then a contact zone is formed, which
is called a Schottky barrier. As such a contact is formed, then the electron charge will move from
the metal to the semiconductor. As a result, the semiconductor becomes negatively charged with
respect to the metal. This transfer process continues until it reaches a state of equilibrium (i.e., the
rate of electron transfer becomes the same in both directions, as the Fermi level is the same in both
of these materials). There will be a space charge in the region of the semiconductor in contact with
the metal. As a consequence, a band bending is observed (Figure 3.6).
Reverse band-bending is observed when a p-type semiconductor is in contact with the metal
(Figure 3.7).

n-SC Metal

CB
EF

VB
FIGURE 3.6  Band bending in n-SC/metal contact.
20 Solar Energy Conversion and Storage

p-SC Metal

CB

EF

VB

FIGURE 3.7  Band bending in p-SC/metal contact.

There will be a potential difference between the semiconductor and the metal as the charge car-
riers will be moved in opposite directions. If the circuit is closed by an external load, some current
will pass, and the strength of this current is proportional to light intensity.
Such a contact zone is also formed if two different kinds of semiconductor (n-type and p-type)
are kept in contact. This is called a case of n-p junction, and the contact zone is called an n-p barrier
(Figure 3.8). Illumination of this combination also will produce the current in an external circuit,
which is proportional to the intensity of light used. These cells are known as Schottky-type cells if
there is contact between the semiconductor and the metal; they are known as n-p-type cells where
electrodes in contact are n-doped and p-doped semiconductors.
If these is an n-type semiconductor and an electrolyte containing a redox couple (An+ An+1) is in
contact, then the Fermi level of the semiconductor will be at a higher energy than the redox potential
of the couple (Figure 3.9).
If these two phases are brought into contact with each other, then the system will try to attain
equilibrium, and the Fermi energies of two phases will become equal. There will be a transfer of
electrons from the semiconductor to the electrolyte resulting in adjustment of the Fermi level. As
the conductance is electronic in nature in the solid phase (semiconductor) but it is ionic in nature

n-SC p-SC
CB

CB
EF
VB

VB
FIGURE 3.8  Band bending in n-SC/p-SC contact.

CB
EC
EF
A+/A2+

EV
VB
Electrolyte
n-Type Vacuum
FIGURE 3.9  n-Type semiconductor/redox couple in vacuum.
Basic Photoelectrochemistry 21

in a solution (electrolyte), electrons cannot penetrate the interface, and transfer of the charge has to
take place by a reduction process:

A(n+1) + e− → A n+ (3.1)

This electron transfer process produces a positive charge on the semiconductor, and a surface region
is formed with a lower concentration of electrons than in the bulk of the semiconductor. This region
is called the depletion layer or the space charge layer. As a result, there will be a corresponding
increase in the number of electrons in the electrolyte. Thus, a negative countercharge is developed in
the electrolyte. Due to this low concentration of charge carrier in the semiconductor as compared to
that in electrolyte, the thickness of the depletion layer is quite higher in magnitude than the negative
countercharge region in the electrolyte.
An electric field exists in this depletion layer because of this charge transfer. This electric field
will change the electrostatic potential with the distance from the surface. If the potential related to
a reference level is Vs at the surface, then it will have another value VB in the bulk of the semicon-
ductor. The difference between these two potentials, (VS – VB), is called the band bending. This
variation in potential corresponds to the change in energy level as potential and energy are almost
identical quantities for the electron (Figures 3.10 and 3.11).
The width (W) of this depletion layer is approximately equal to

1/2
⎛ 2εε0VB ⎞
W =⎜ ⎟ (3.2)
⎝ eN D ⎠
where e is the elementary charge, ND is the concentration of ionized donor atom, ε is the dielectric
constant of the material, and ε0 is the permittivity of free space. The semiconductor–electrolyte
interface has many properties similar to the metal–semiconductor interface; therefore, it is quite
commonly referred to as a Schottky-type junction.

Space charge Electrolyte


layer
n-Type
Helmholtz layer
FIGURE 3.10  Space charge layer in n-type SC/electrolyte system.

VB Ec,s
CB
Ec,b
A+/A2+
EF

EV,S

Ev,b
VB
EV
n-Type Electrolyte FIGURE 3.11  n-Type semiconductor/electrolyte with a redox couple.
22 Solar Energy Conversion and Storage

EC

A+/A2+

EF
EV
VB

p-Type Vacuum Electrolyte


FIGURE 3.12  p-Type semiconductor/redox couple in vacuum.

In the case of a p-type semiconductor in contact with an electrolyte, the case is reversed. A Fermi
level in this case is quite close to the valence band (Figure 3.12).
In the case of a p-type semiconductor in contact with the electrolyte, the redox level is above that
of the Fermi level of the semiconductor. As a result, electrons will start flowing from the electrolyte
into the semiconductor until the equilibrium is attained. In such a situation, a negative space charge
layer is formed and the energy level is bent upward at equilibrium. This means that the energy level
will be higher in bulk than at the surface. A positive countercharge is formed in the electrolyte close
to the semiconductor. Here, the width of the depletion layer is
1/2
⎛ 2εε0VB ⎞
W =⎜ ⎟ (3.3)
⎝ eN A ⎠
with the only difference that ND is replaced by NA, the concentration of ionized accepter ions (Figure 3.13).
The position of the Fermi level with some reference level of the solution (normally H2/H+ level)
is called the potential of the semiconductor. The Fermi level of the semiconductor can be varied
relative to this reference level by applying some external potential to the semiconductor electrode.
Therefore, a band bending can be changed from value at equilibrium. Whenever, there is no field in
the semiconductor, (VS = VB), then the potential is called the flat-band potential (Vfb). This flat-band
potential decides the position of the conduction band of the n-type semiconductor and valence band
of the p-type semiconductor, approximately.
The flat-band potential can be determined by measuring the differential capacitance (Cdl) of the
space charge-transfer layer as a function of the potential of the semiconductor electrode. According
to the Mott–Schottky equation (Myamlin and Pleskov 1967),

⎛ kT ⎞
Cdl−2 = (2εε0 eN D ) ⎜V −V fb − ⎟ (3.4)
⎝ e ⎠
where k is the Boltzmann constant, T is the temperature, and Cdl is the differential capacity.

Space charge Helmholtz Electrolyte


layer layer
FIGURE 3.13  Space charge layer in a p-type semiconductor/electro-
p-Type
lyte system.
Basic Photoelectrochemistry 23

A plot of  1/Cdl2   v/s potential (V) of the semiconductor gives a straight line. This line cuts the
potential axis near the flat-band potential (kT/e ≈ 0.025 V at room temperature). The slope of this
line is determined by the concentration of the doping atom (ND in the case of an n-type material
or NA in the case of a p-type material). The measured capacitance is a complex quantity, because
it is a sum of various contributing capacitances of the system, each belonging to the space charge.
Tomkiewicz (1979) has discussed the measurement of the space charge layer capacitance.
It has been observed that the flat-band potential is found to change with the pH of the solution
for some semiconductors (mostly oxides). This ability of the semiconductor to change with the pH
may be explained based on an assumption. If the surface of the semiconductor acts as an acid–base
couple, which can either accept or donate protons, then

M − OH = M − O− + H+ (3.5)

The potential difference between the surface of the semiconductor and the bulk of the electrolyte,
which is approximately the potential drop of the Helmholtz layer (assuming that the contribution of
Gouy layer is negligible) will be

ΔVH = constant – 0.059 pH (3.6)

When there is a vacuum between the semiconductor and the electrolyte, then the Fermi energy level
of the semiconductor is termed EFvac (Figure 3.14).
As the electrolyte and semiconductor approach each other, a field is induced, which causes
a change of the vacuum layer level in the region between these two phases (Figure 3.15). When
this field is completely developed after the Helmholtz layer is completed, then the potential drop

CB E
Ec,b
VB
Ec,s
A+/A2+
EF

VB

Ev,b
EV,S

p-Type Electrolyte
FIGURE 3.14  p-Type semiconductor/electrolyte with a redox couple.

EC –4.5 Vacuum
level
X
CB EV.C
EF
H2/H+
0
EV
VB
n-Type Vacuum Electrolyte
(VS NHE)
FIGURE 3.15  n-Type semiconductor in vacuum.
24 Solar Energy Conversion and Storage

DVH
X
DVH

VB
CB
H2/H+
EF

V
VB
Electrolyte
n-Type
FIGURE 3.16  Effect of pH on potential drop in n-type semiconduc-
Helmholtz layer
tor/electrolyte with H2/H+ redox couple (at lower pH).

ΔVH = VS – VH must be considered in relating the energy levels of the two phases to each other. The
flat-band potential is

V fb = −(EFvac /e + 4.5 V − ΔVH ) (3.7)

A change in pH also affects the electrode potential of the H2/H+ couple (Figure 3.16) as

V H 2 /H+ = –0.059 pH +V 0 H 2 /H+ (3.8)


( ) ( )
At equilibrium, the Fermi energy (i.e., the potential of the semiconductor electrode) will show simi-
lar pH dependence. It means that a band bending is independent of pH.

3.2 PHOTOELECTROCHEMISTRY
If an n-type semiconductor is in contact with an electrolyte, and it is exposed to a light of energy
greater than or equal to Eg (bandgap of a semiconductor), then the photons are absorbed by the
semiconductor resulting in excitation of electrons from the valence band to the conduction band
(Figure 3.17). This creates electron–hole pairs. This electron–hole pair may recombine; therefore,
it must be quickly separated into a free electron and a free hole. This separation is achieved in the

∆VH
X

VB CB
H2/H+
EF

V
VB
FIGURE 3.17  Effect of pH on potential drop in n-type semiconductor/ Electrolyte
n-Type
electrolyte with H2/H+ redox couple (at higher pH).
Basic Photoelectrochemistry 25

CB EC
EF*
∆V *F
A+/A2+
EF

EV

FIGURE 3.18  Effect on the band bending and Fermi level on excitation
VB
of n-type semiconductor/electrolyte system.

space charge layer, where a strong electric field is effective. The majority carriers (electrons) will
pass to the bulk of the semiconductor as the minority carriers (holes) will be driven to the surface.
If this excitation takes place in the region behind the depletion layer (i.e., the diffusion region), the
electron–hole pairs will recombine unless the holes diffuse away into the depletion layer, as there is
no field in the diffusion region. The excitation of a semiconductor in the depletion layer is an excel-
lent means of separating the negative and positive charges quickly, and it is necessary for a system
to be a good working system.
According to an assumption, no reaction at the surface can exhaust the holes; hence, a photopo-
tential develops in the semiconductor, decreasing the band bending. As a consequence, the Fermi
level will change toward its flat-band position (Figure 3.18).
This photopotential (VPhP) is

(EF* − EF ) ΔE *
VPhP = − = − F (3.9)
e e

where EF is the Fermi level at equilibrium, and EF* is the Fermi level under illumination. The pho-
topotential VPhP can also be shown as

⎛ kT ⎞
VPhP ~ ⎜ ⎟ ln I (3.10)
⎝ e ⎠

where I represents the light intensity. As the value of

EF* (max)
= −VFb − 4.5 V
e
the maximum photovoltage obtained will be (Gerischer 1975)

EF
max
VPhP = Vfb + 4.5V + (3.11)
e
But in practice, this much photopotential is not obtained, maybe because of its logarithmic depen-
dence on intensity. The flow (J) of minority carriers to the surface is the sum of the flows from the
depletion layer (Jdl) and the diffusion region (Jdiff ). This flow is given by (Gartner 1959)

⎡1− exp (−α λW ) ⎤


J = J dl + J diff = eI λ ⎢ ⎥ (3.12)
⎣ 1+α λ L ⎦
26 Solar Energy Conversion and Storage

where the wavelength of monochromatic radiation is λ, and its intensity is Iλ. If αλ is the absorption
coefficient of the semiconductor, then the diffusion length (L) of the minority carrier is

L = (Dτ )1/2 (3.13)

where D is the diffusion coefficient, and τ is the lifetime of minority carriers.


Whether the minority carriers reaching the surface will lead to a photocurrent depends on the
probability of charge transfer (St) and the surface in relation to the probability of recombination (Sr)
with surface states. Such a relation for the generated photocurrent will be (Wilson 1977)

⎛ St ⎞
⎜ ⎟ elλ [1− exp(−α λW )]
⎝S +S ⎠
iPhP = t r (3.14)
1− α λW

The charge transfer should preferably involve reaction with species in the electrolyte, but it may also
involve a corrosion reaction leading to a photocurrent.

3.3 PHOTOCATALYSIS
Photocatalysis is an emerging and promising technology that has varied applications. The term pho-
tocatalysis creates confusion, as it is a combination of two terms: photo, meaning “light,” and cataly-
sis, meaning “to affect the rate of a reaction.” It appears as if photocatalysis means light-catalyzed
reaction, but that is not the case. This term has long been debated, but the term photocatalysis is still
reserved for a chemical reaction in the presence of light and a semiconductor.
Photocatalysis has been classified into two types:

• Homogeneous
• Heterogeneous

Homogeneous photocatalysis involves a photocatalytic reaction, where a substance and the semi-
conductor are in the same phases. The best examples are dyes, coordinatation compounds, and
so on. Heterogeneous photocatalysis involves the semiconductor and substrate in different phases.
Common examples are insoluble semiconducting chalcogenides in binary, ternary, and sometimes,
even quaternary forms.
The concept of photocatalysis (Figure 3.19) is based on photoelectrochemistry, but with a difference
that here the particles of the semiconductor act as individual photoactive units. When the semiconduc-
tor is exposed to a suitable light radiation with wavelengths corresponding to equal or more than its
bandgap, then an electron will be excited from its valence band to the conduction band, thus leaving
behind a hole in the valence band. This electron can be utilized to reduce a substrate, or the hole can
be used for oxidizing any substance, of course, depending upon the redox levels of the substrate.
It is not normally possible to have an oxidizing as well as a reducing environment simultane-
ously in the same system. The beauty of photocatalysis is that it provides both of these environments
simultaneously in the same reaction medium. However, reduction and/or oxidation may take place
depending upon a situation.
In all, there are four possibilities (Figure 3.20) depending upon the levels of conduction band,
valence band, and redox levels of the substrate:

• If the reduction level of the substrate is below the CB of the semiconductor and the oxida-
tion level is below its VB, then transfer of electrons from the semiconductor to the substrate
becomes easier, resulting in reduction of the substrate.
Basic Photoelectrochemistry 27

Reduced

Excited
Light source e−
Reduction
(hν) (light)
A + e−→ A−•
Conduction band
e−
Acceptor

Energy Bandgap

Donor
+
Valence band h
Oxidation
Hole D + h+→ D+•

Oxidized

FIGURE 3.19  Photocatalysis.

Conduction
band –
e

h+
Valence
band
Reduction Oxidation Oxidation No
and reduction reaction

FIGURE 3.20  Various probabilities of reactions.

• If the reduction level is higher than the CB of the semiconductor, and the oxidation level
of the substrate is higher than the VB of the substrate, then the electron can be easily
transferred from substrate to semiconductor. In other words, the hole is transferred from
semiconductor to substrate, resulting in oxidation of the substrate.
• If the reduction level of the semiconductor is lower than the CB of the semiconductor and
the oxidation level is higher than the VB of semiconductor, then it appears as if the CB and
VB of the semiconductor saddle over the redox levels of the substrate. In such cases, the
electron and hole can be transferred from the semiconductor to the substrate. Here, both
oxidation and reduction will take place.
• If the reduction level of the substrate is higher than the CB of the semiconductor, and VB
is lower than the oxidation level of the semiconductor, then no reaction will take place.
28 Solar Energy Conversion and Storage

Photocatalysis has provided some interesting applications like wastewater treatment, deodorization,
self-cleaning glasses, disinfection, synthesis of energy-rich materials, photogeneration of hydrogen,
antifogging reduction of CO2, solar energy conversion and storage, and so on.

REFERENCES
Gartner, W. W. 1959. Depletion-layer photoeffects in semiconductors. Phys. Rev. 116: 84–87.
Gerischer, H. 1975. Electrochemical photo and solar cells. Principles and some experiments. J. Electroanal.
Chem. 58: 263–274.
Myamlin, V. A., and Pleskov, V. Yu. 1967. Electrochemistry of Semiconductors. Plenum Press, New York.
Tomkiewicz, M. 1979. Relaxation spectrum analysis of semiconductor-electrolyte interface-TiO2. J. Electrochem.
Soc. 126: 2220–2225.
Wilson, R. H. 1977. A model for the current-voltage curve of photoexcited semiconductor electrodes. J. Appl.
Phys. 48: 4292–4297.
4 Photoelectrochemical Cells
Dipti Soni, Priya Parsoya, Basant K. Menariya,
Ritu Vyas, and Rakshit Ameta

CONTENTS
4.1 Introduction............................................................................................................................. 29
4.2 Classification............................................................................................................................ 30
4.2.1 Regenerative Solar Cells.............................................................................................. 31
4.2.2 Photoelectrolytic Solar Cells....................................................................................... 39
4.2.3 Photocatalytic Cells..................................................................................................... 42
4.3 PEC Cells as Storage Cells...................................................................................................... 47
References......................................................................................................................................... 48

4.1 INTRODUCTION
Photoelectrochemical (PEC) cells are the most efficient cells for converting solar energy into a more
useful form of energy. These devices are quite simple to construct, and often consist of a photoactive
semiconductor electrode (either n- or p-type) and a metal counterelectrode. Both of these electrodes
are immersed in a suitable redox electrolyte. The PEC cells use light to carry out a chemical reac-
tion for converting light to chemical energy. They have a solid–liquid interface, whereas photovol-
taic (PV) solar cells have a solid–solid interface.
The commercial use of a PEC solar cell depends on its conversion efficiency and stability. Various
efforts have been made to make PEC cells more efficient, such as electrolyte modification, surface
modification of the semiconductors, photoetching of layered semiconductors, semiconductor septum–
based PEC solar cells, and so on. The dream is to capture the energy that is freely available from
sunlight and turn it into electric power.
The PEC cells based on III–V semiconductor electrodes have achieved high solar power con-
version efficiencies in regenerative cells and in photoelectrolytic production of hydrogen. Miller
(1984) discussed the corrosion chemistry associated with charge transfer at these interfaces, the
influence of film formation, and the consequences for both photoanodic and photocathodic cells.
Single-bandgap semiconductors in PEC cells have lower values (up to 16%) of energy conversion
than the multiple bandgap cells that have significantly higher conversion efficiencies (Licht, 2001;
Licht et al. 1998b,c,d).
Energy production by these PEC processes has also been reviewed by Memming (1978). Bhavani
et al. (1986) studied the reactions and PECs from the standpoint of energy conversion efficiency and
the possibility of energy storage. McEvoy (2005) demonstrated that in a PEC, injected electrons
facilitate a current in an external circuit, returning to the redox electrolyte through a cathode that
is in contact with it. The uncharged ground state of the dye is restored by electron transfer from the
redox system, which completes the circuit and provides a regenerative cycle comparable with other
photovoltaic devices.
French scientist Edmond Becquerel (1839) noticed a photovoltaic effect when he immersed a sil-
ver electrode in a chloride electrolyte and it was illuminated. Fujushima and Honda (1972) worked
on illuminated semiconducting TiO2, which leads to photooxidation of water to oxygen. They pho-
toelectrolyzed water into H2 and O2 by using such a system. This process results in the conversion of
sunlight to stored chemical energy. Metal oxide semiconductors like TiO2, SrTiO3, WO3, and so on,

29
30 Solar Energy Conversion and Storage

are very stable and have large bandgaps, which converts only a small fraction of the solar spectrum
into electrical or chemical energy. It is challenging to find a semiconductor material with a small
bandgap (1.1–1.5 eV) for efficiently converting sunlight to usable energy.
The semiconductor–liquid-junction-based PEC solar cell consists of a photoactive semiconductor
electrode immersed in electrolytic solution containing suitable redox couple and counterelectrode,
which can be a metal or semiconductor. Irradiation of the semiconductor–electrolyte junction with
light of hν > Eg (Eg is the bandgap of SC) results in the generation and separation of charge carriers.
The majority of carriers are electrons in an n-type semiconductor, which move to counterelectrode
through an external circuit and take part in a counter reaction. Holes are the minority charge carri-
ers, which in turn migrate to electrolytes and participate in electrochemical reactions.

4.2 CLASSIFICATION
The PEC cells can be classified in two major categories on the basis of change in Gibbs free energy:

1. Regenerative PEC solar cells with ΔG = 0. Here, the photoenergy is converted into electric
energy (Figure 4.1).
2. Photoelectrosynthetic cells with ΔG ≠ 0. Here, the photoenergy is used to affect chemical
reactions, with nonzero free energy change in the electrolyte. These cells can be further
classified into two types of cells:
a. Photoelectrolytic cells with ΔG  >  0, where the photoenergy is stored as chemical
energy in endergonic reactions (e.g., H2O → H2 + ½O2 ) (Figure 4.2).
b. Photocatalytic cells with ΔG < 0, where photoenergy provides activation energy for
exergonic reactions (e.g., N2 + 3H2 → 2NH3) (Figure 4.3).

e e

e e
e
O O
R R
ΔG = 0
e

n-SC Soln M p-SC Soln M

FIGURE 4.1  Regenerative solar cells.

e e

e e
O
´ O
R ´ R
ΔG > 0
O O
R R
e

n-SC Soln M p-SC Soln M

FIGURE 4.2  Photoelectrolytic solar cells.


Photoelectrochemical Cells 31

e e

O
R e
e O
ΔG < 0
O ´ R
R ´ O
R

n-SC Soln M p-SC Soln M

FIGURE 4.3  Photocatalytic solar cells.

Three electrical parameters determine the conversion efficiency of a liquid-junction PEC cell:

• Open-circuit voltage (Voc)


• Short-circuit photocurrent (isc)
• The fill factor (FF)

An efficient PEC device can be constructed by optimizing these three quantities.


Solar-to-electrical conversion efficiency (η) in a semiconductor/liquid-junction solar cell is

⎡ (i ×V )max ⎤
η=⎢ ⎥⎦ ×100 (4.1)
⎣ P
where (i × V)max is the maximum output power of the solar cell, and P is the optical power input.
Fill factor (FF) of a PEC solar cell is

imax ×Vmax
FF = (4.2)
isc ×Voc

The deviation of fill factor (normally less than unity) indicates the extent of departure from the ideal
i-V behavior.
The maximum η for operation in regenerative (light to electricity) mode is 25%–30% for single
electrode–based PEC cells employing semiconductors with bandgaps of 1.2–1.5 eV. If the optimum
bandgap for photoelectrolysis is considered to be 1.8 eV, then the theoretical maximum efficiency
is estimated to be 25%, while for double photoelectrode–based (p-n) water photoelectrolysis cells,
it is almost 45%.

4.2.1  Regenerative Solar Cells


Regenerative PEC solar cells are used to convert solar energy to electricity. Regenerative PEC solar
cells, which are based on a narrow-bandgap semiconductor and a redox couple, convert optical
energy into electrical energy without bringing about any change in the free energy of the redox
electrolyte (ΔG = 0). The electrochemical reaction occurring at the counterelectrode is opposite to
the photoassisted reaction occurring at the semiconductor working electrode. Thus, they are also
called electrochemical photovoltaic cells. Electrochemical photovoltaic cells have the following
advantages over solid photovoltaic cells:

• They are not sensitive to the defects in semiconductors.


• The solid–liquid junction is easy to form, and the production price will be much lower.
32 Solar Energy Conversion and Storage

• Direct energy transfer from photons to chemical energy is possible. Unlike conventional
solid-state photovoltaic cells, the potential of the working electrode can be varied with
respect to the reference electrode by means of an external voltage source connected
between the working electrode and the counterelectrode.

Doubly β-functionalized porphyrin sensitizers were prepared by Park et al. (2008) to study the
photoelectrochemical properties of dye-sensitized nanocrystalline-TiO2 solar cells. These porphy-
rin sensitizers were functionalized at meso- and β-positions with different carboxylic acid groups,
which were then employed to investigate electronic and photovoltaic properties. Multiple pathways
through olefinic side chains at two β-positions enhance the overall electron injection efficiency,
and the moderate distance between the porphyrin sensitizer and the TiO2 semiconductor layer will
retard the charge recombination processes. Consequently, these combined effects give rise to higher
photovoltaic efficiency in photovoltaic regenerative solar cells.
Kohl and Bard (1979) constructed regenerative photoelectrochemical cells by using single-crystal
n-GaAs in acetonitrile solutions. Solution redox couples (anthraquinone, p-benzoquinone, dimethyl
ferrocene, ferrocene, hydroxymethyl ferrocene, and tetramethyl-p-phenylenediamine) were reduced
at a Pt counterelectrode and photooxidized at the semiconductor electrode converting light directly
into electrical energy. A power conversion efficiency of 14% was obtained for the n-GaAs electrode
in a ferrocene–ferricenium acetonitrile solution at a radiant intensity of 0.52 mW cm–2.
Tributsch (1980) studied the photoelectrochemical properties of semiconducting layer-type
disulfides and diselenides of transition metals belonging to groups IV, VI, and VIII, which have
energy gaps ranging between 1 and 2 eV. The compounds of Mo and W can be used as stable elec-
trodes for regenerative and fuel-producing solar cells. Neumann-Spallart and Kalyanasundaram
(1981) fabricated polycrystalline CdS electrodes from CdS powders by coating a thin film of a paste
formed with an aqueous solution of a surfactant and ZnCl2 on a Ti substrate and sintering at 670°C
in argon. Light conversion efficiencies of these electrodes have been tested in two regenerative solar
cells and were found to range from 0.37% to 4.4%.
Singh et al. (1981) reported the photoelectrochemical behavior of n-GaAs electrodes used in
regenerative PEC cells. Both single-crystal and polycrystalline n-GaAs electrodes were used, and
it was concluded that this PEC system was capable of generating high open-circuit potentials. Kline
et al. (1981) prepared and used single crystals of n-WSe2 and n-MoSe2 as the photoanodes in a regen-
erative PEC cell with iodide/triiodide electrolyte. The conversion efficiencies observed were 10.2%
and 9.4% on selected crystals of WSe2 and MoSe2, respectively.
Kline et al. (1982) studied the PEC behavior of synthetic crystals of WS2, MoS2, and crystals with
mixed-metal and chalcogen composition and compared that with the behavior of MoSe2 and WSe2.
The composition and stoichiometry of the crystals and the composition of the electrolyte were var-
ied, and the behavior of the materials in a regenerative liquid-junction solar cell was observed. The
quantum yields and sunlight-to-electricity conversion efficiencies were also measured.
Catalytically modified n-silicon/indium tin oxide anodes were used in PEC generation of chlo-
rine by Thompson et al. (1982). Silicon photoanodes showed stable and efficient behavior Cl2/Cl
with aqueous LiCl electrolytes in PEC cells. Regenerative devices using the Cl2/Cl− redox coupled
with approximately 3% optical-to-electrical conversion efficiency have been reported. Unmodified
n-Si/ITO electrodes show minimal electro- and photocatalytic activity for the generation of Cl2.
Ang et al. (1983) made a two-photoelectrode regenerative PEC cell, which showed a combined
photovoltage in excess of 1 V. The cell consists of n-CdSe as the photoanode and p-InP as the photo-
cathode in an aqueous sulfide/polysulfide electrolyte. The photo potentials generated were sufficient
for the external change of a number of redox battery couples. Rajeshwar et al. (1983) used polymer-
coated n-GaAs photoanodes with aqueous electrolyte in regenerative PEC cells. The n-GaAs pho-
toanodes were coated with films of polystyrene pendant [Ru(bpy)3]2+ complex, and it was called a
PSt-bpy-Ru complex. The PSt-bpy-Ru-coated n-GaAs photoanodes were tested in aqueous redox
electrolytes in regenerative PEC cells.
Photoelectrochemical Cells 33

The behavior of regenerative PEC cells of large-grained polycrystalline n-CuInSe2 films was studied
by Bicelli et al. (1985). These films were synthesized by a new technique of depositing the semicon-
ducting material on thin metal layers and then keeping it at a temperature lower than its melting point.
Scanning electron microscopy was used to characterize the surface morphology of the films; sys-
tematic scanning laser spot analysis was employed to examine the output power characteristics when
in contact with different redox couples. Keita and Nadjo (1984) used a single redox couple, sodium
9,10-anthraquinone-2,6-disulfonate/sodium 9,10-anthrahydroquinone-2,6-disulfonate (AQ/AQH2) to
illustrate various aspects of PEC cells, particularly regenerative photocells with p-type Si and p-type
WSe2 photocathodes and synthetic photocells. The conversion efficiencies were in the range of 2% for
p-type Si electrode and 10% for layer-type materials at 632.8 nm light.
Osaka et al. (1985) studied the iron oxide/n-Si heterojunction electrode as a photoanode for a
regenerative PEC cell. The effects of modifying the top layer of the electrode with Pd or RuO2 were
also observed. The photocurrent at the heterojunction electrode was generated by the holes, which
were photoexcited in both iron oxide and n-Si. The addition of Pd or RuO2 on the heterojunction
electrode surface increased the optical-to-electrical conversion efficiency. The efficiencies for a sta-
ble working PEC cell were found to be 1.34% and 1.60% for a Pd- and a RuO2-modified electrode,
respectively. This efficiency was measured in a 0.2 M KOH solution containing 0.2 M K4[Fe(CN)6]
and 0.01 M K3[Fe(CN)6] at an intensity of 55 mW cm−2. The photoanode was made highly stable by
the use of iron oxide, as compared to an electrode such as RuO2/n-Si.
Cattarin et al. (1986) studied the photoelectrochemical behavior of copper mono- and disubstituted
para-diethynylbenzene, deposited chemically and electrochemically as thin films onto copper elec-
trodes in 0.1 M LiOH. These materials act as p-type semiconductors with bandgaps of 2.2 and 2.1 eV,
respectively, in a rectifying liquid junction with a flat-band potential of 0.3 V versus SCE. High photo-
currents (8.0 mA cm–2) and photovoltages (0.5 V) were observed by the disubstituted compound under
oxygen and 35 mW cm–2 white light. The monochromatic quantum yields of 12%–13% were obtained.
These results displayed a remarkable improvement with respect to evaluated copper acetylide systems.
The solar-to-electricity conversion efficiency of a regenerative PEC cell was determined by
Carlsson and Holmström (1986). The analysis was based on two types of laboratory measurements:
(1) photocurrent as a function of photoelectrode potential at constant wavelength of illuminating
light and (2) photocurrent as a function of wavelength at constant potential.
Guay et al. (1987) obtained high short-circuit photocurrents with phthalocyanine in a regen-
erative PEC cell. Purified phthalocyanine-AlCl was synthesized, which generated photocathodic
short-circuit photocurrents in the milliampere per square centimeter (mA cm–2) range under white
light irradiation.
Ramaraj and Natarajan (1989) synthesized and characterized macromolecular-bound phenosaf-
ranine dyes. Photoelectrochemical investigations of these macromolecular phenosafranine dyes
showed different behaviors depending on the macromolecule. Cathodic behavior with reference to
an inert electrode was observed when the electrode was coated with a film of poly(acrylamidometh
ylphenosafranine-co-methylolacrylamide), while an electrode coated with a film of poly(acrylamid
omethylphenosafranine-co-methylolacrylamide-co-vinylpyridine) exhibited anodic polarity. These
polymeric phenosafranine-coated electrodes were used to operate a water-splitting regenerative cell.
Ladouceur et al. (1990) prepared two substoichiometric tungsten oxide films by plasma spray of
WO3 powder on Ti substrates. The films were 40 ± 20 μm thick and yellow (WO2.99) or dark blue
(WO2.97) in color. The yellow films have been used in regenerative cells by using O2/H2O redox
couple at pH 2.0. The WO2.99 films have been analyzed for NH3 photoproduction. Chloro- and bro-
moaluminum phthalocyanines were studied in cells using I3−/I− as the redox system.
Rothenberger et al. (1992) synthesized nanostructured TiO2 films by sintering 15 nm diameter
colloidal anatase particles on a conducting glass support. They optically determined the flat-band
potential of colloidal titanium dioxide films. The ability to measure flat-band potentials in such
porous electrodes helps in the optimization of interfacial charge-transfer processes and in the design
of efficient regenerative PEC cells.
34 Solar Energy Conversion and Storage

Redmond et al. (1994) prepared transparent nanocrystalline ZnO films on a conducting glass substrate
by sol-gel techniques. They described the sensitization by adsorption of a ruthenium-based complex and
subsequent incorporation as the light-harvesting unit in a regenerative PEC cell. The resulting device had
a monochromatic incident photon-to-current conversion efficiency of 13% at 520 nm.
The effect of aqueous polyselenide solution modification on the photoelectrochemical behavior
of an n-GaAs/aqueous polyselenide PEC cell was described by Forouzan and Licht (1995). Solution
modification was achieved by changing the ratio of dissolved selenium to selenide and pH control. It
was concluded that PEC photocurrent, photovoltage, and fill factor were affected by the distribution
of hydroselenide, selenide, and polyselenide in solution. An optimum pH ([KOH] = 1–2 M) pro-
vides sufficient concentrations of Se2− to improve PEC response. The photopotential was enhanced
up to 50 mV in high-Se° electrolytes.
Matthews et al. (1996) calculated the photocurrent-potential characteristics for regenerative, sensi-
tized semiconductor electrodes. The steady-state anodic photocurrent for a sensitized semiconductor
electrode has been reported, taking into account the rates of light absorption by the sensitizer, electron
injection from the excited state of the sensitizer to the conduction band of the semiconductor, decay of
the excited state, and reductive regeneration of the sensitizer by the redox electrolyte. In this model,
the rate of recombination between the conduction band electron and the oxidized sensitizer and the
reactions between the excited state and the redox couple have been assumed to be negligible.
Papageorgiou et al. (1996) investigated the physical–electrochemical properties of methylhexy-
limidazolium iodide (MHImI) and its mixtures with organic solvents, such as n-methyloxazolidi-
none and acetonitrile, and with other lower-viscosity molten salts, such as methyl-butylimidazolium
triflate. Roušar et al. (1996) worked on the optimization of physical and geometrical parameters
of an electrochemical photovoltaic regenerative solar cell with current leads located on opposite
sides of a cell unit. The dependence of local current density on the length coordinate was expressed
using dimensionless quantities, and a linear polarization curve was assumed. The optimization of
the output on the ratio of the active surface to the total surface of the cell was explained, and the
construction of a cell with line current collectors was also discussed.
The photoelectrochemical measurements of the modified TiO2 electrodes in regenerative solar
cells were reported by Heimer et al. (1996). The coordination compounds of the type Ru(dmb)2(LL)
(PF6)2, where dmb is 4,4′-(CH3)2-2,2′-bipyridine and LL is 4-(CH3)-4′-(COOH)-2,2′-bipyridine,
or 4-(CH3)-4′-((CH2)3COOH)-2,2′-bipyridine, or 4-(CH3)-4′-((CH2)3- COCH2COOC2H5)-2,2′-
bipyridine, were prepared for attachment to the TiO2 surface. The optical and redox properties of
these compounds in dichloromethane solution were investigated. The Langmuir adsorption iso-
therm model was used to analyze the binding to porous nanostructured TiO2 films.
Garcia et al. (1998) used 4-phenylpyridine as ancillary ligand in ruthenium(II) polypyridyl
complexes for sensitization of n-type TiO2 electrodes. Two types of molecular sensitizers, cis-
[(dcbH2)2Ru(ppy)2]2+ and cis-[(dcbH2)2Ru(ppy) (H2O)]2+ (dcbH2 is 4,4′-(CO2H)2-2,2′-bipyridine; ppy
is 4-phenylpyridine), have been prepared. One coordinated ppy, cis[(dcbH2)2Ru(ppy)(H2O)]2+, has a
higher incident photon to current efficiency (IPCE) value than the corresponding derivative with two
coordinated ppy, cis-[(dcbH2)2Ru(ppy)2]2+.
Falaras et al. (1998) synthesized and characterized Ru(PPh3)2(dcbipy)Cl2 (PPh3 is triphenylphos-
phine; dcbipy is 2,2′-bipyridyl-4,4′-dicarboxylate) for efficient photosensitization of titanium oxide
in wet regenerative PEC cells. The broad bands in the visible spectrum as well as the reversibility of
the redox couple RuIII–RuIII make this complex potentially beneficial for the photosensitization pro-
cess. Vlachopoulos et al. (1988) used high surface area polycrystalline anatase films together with
tris(2,2′-bipyridyl-4,4′-dicarboxylate)ruthenium(II), RuL34−, as a sensitizer, and achieved efficient
visible light energy conversion to electric current. In the presence of iodide as an electron donor,
incident photon-to-current conversion efficiencies of 73% have been obtained at the λmax of the dye.
Bromide is oxidized under the same conditions with an efficiency of 56%. A regenerative cell based
on the Br2/Br− redox system gives a monochromatic light-to-power conversion efficiency of 12%
with a fill factor of 0.74.
Photoelectrochemical Cells 35

A low power fullerene PEC solar cell, utilizing a regenerative polyiodide and ferri/ferrocyanide
redox couple was studied by Licht et al. (1998a). They have demonstrated the photoelectrochemistry
of illuminated and immersed single-crystal C60, which has been shown to drive oxidation of several
solution-phase redox couples. Utilization of a PEC solid–liquid junction, rather than a solid-state
photovoltaic junction, resulted in improvement in the observed photocurrent. The spectral response
and current–voltage behaviors in several electrolytes were also observed. A higher temperature and
C60 oxygen depletion increase the photocurrent of fullerene PEC solar cells. Fullerene/iodide elec-
trolyte PEC cells consisting of intrinsic single-crystal C60 in aqueous 3 M KI, 0.01 M I2, or 0.1 M
tetrabutyl ammonium iodide, 0.3 M LiClO4 in acetonitrile solution, drive regenerative photoinduced
iodide oxidation. The photocurrent was increased by an order of magnitude 6.4  μA  cm−2 under
100 mW cm−2 illumination by an increase in the aqueous cell temperature from 24°C to 82°C. In a
similar way, photocurrent was increased by O2 depletion pretreatment (24 h at 400°C in Ar) (Licht
et al. 1998e).
Argazzi et al. (1998) synthesized new Ru(dcbH)(dcbH2)(L) sensitizers, where L is diethyldithiocar-
bamate, dibenzyldithiocarbamate, or pyrrolidinedithiocarbamate; dcbH is 4-(COOH)-4′-(COO –)-2,2′-
bipyridine; and dcbH2 is 4,4′-(COOH)2-2,2′-bipyridine. These have been used in nanocrystalline TiO2
films for light to electrical energy conversion in regenerative PEC cells with I−/I3− acetonitrile electro-
lyte. Photophysical measurements show that the high photocurrent observed for cis-Ru(dcb)2(NCS)2/
TiO2 was due to efficient and rapid iodide oxidation.
The cis-Os(dcb)2(CN)2/TiO2 PEC cells have the limiting role in iodide oxidation (Alebbi et al.
1998). In situ time-resolved diffuse reflectance measurements showed that a slow iodide oxidation
rate is responsible for the low photocurrent efficiency of cis-Os(dcb)2(CN)2/TiO2.
Yohannes and Inganäs (1999) studied solid-state PEC cells containing a conjugated polymer,
poly[3-(4-octylphenyl)-2,2′-bithiophene], film as a photoactive electrode, a solid polymer electrolyte,
poly[oxymethylene-oligo(oxyethylene)] complexed with redox couple, and a counterelectrode. The
short-circuit current and open-circuit voltage generated with white light illumination at approximately
one sun were 0.4 μA cm−2 and 240 mV, respectively. Two redox couples, I3−/I− and Eu2+/3+, were used.
The active junction between the conjugated polymer and the polymer electrolyte was responsible for
photocurrent generation.
Cation-controlled interfacial charge injection in sensitized nanocrystalline TiO2 was done by
Kelly et al. (1999). The complex of Ru(deeb)(bpy)2(PF6)2, where bpy is 2,2′-bipyridine and deeb is
4,4′-(COOEt)2-2,2′-bipyridine, was anchored to nanocrystalline TiO2 (anatase) or ZrO2 films. Long-
lived metal-to-ligand charge-transfer (MLCT) excited states were observed in acetonitrile (or 0.1 M
tetrabutylammonium perchlorate) on both TiO2 and ZrO2. The addition of Li+ increases both the
efficiency and long wavelength sensitivity of the cell.
Garcia and Iha (2001) employed [(dcbH2)2 RuLL′], where dcbH2 is 4,4′-(CO2H)2-2,2′-bipyridine
and L,L′ is substituted pyridines, as nanocrystalline TiO2 sensitizers in PEC solar cells. This regen-
erative solar cell consists of a transparent conductive oxide (TCO) glass with a dye-sensitized TiO2
semiconductor film as a photoanode, I2/LiI solution in acetonitrile as a redox mediator, and a trans-
parent Pt-coated TCO glass as a counterelectrode. In this cell, 50% IPCE was achieved at 400 and
550 nm.
Nanocrystalline titanium dioxide electrodes in regenerative PEC cells were sensitized by a series of
platinum-based sensitizers of the general type Pt(NN)(SS), where NN is 4,4′-dicarboxy-2,2′-bipyridine
(dcbpy) or 4,7-dicarboxy-1,10-phenanthroline (dcphen), and SS is ethyl-2-cyano-3,3-dimercaptoacry-
late (ecda), quinoxaline-2,3-dithiolate (qdt), 1,2-benzenedithiolate (bdt), or 3,4-toluenedithiolate (tdt)
(Islam et al. 2001a).
Islam et al. (2001b) prepared a new series of ruthenium(II) polypyridyl sensitizers with strong
electron–donating dithiolate ligands Ru(dcbpy)2(L) and Ru(dcphen)2(L), where L is quinoxaline-
2,3-dithiolate (qdt) or ethyl-2-cyano-3,3-dimercaptoacrylate (ecda) or 1,2-benzenedithiolate (bdt)
or 3,4-toluenedithiolate (tdt); dcbpy is 4,4′-dicarboxy-2,2′-bipyridine; and dcphen is 4,7-dicarboxy-
1,10-phenanthroline, for sensitization of nanocrystalline TiO2 electrodes. These complexes show
36 Solar Energy Conversion and Storage

different sensitization to TiO2 electrodes with increasing activity in the sequence (L  =  tdt, bdt,
ecda, qdt) in regenerative PEC cells with I−/I3− acetonitrile electrolyte. Both Ru(dcbpy)2(qdt) and
Ru(dcphen)2(qdt) showed overall cell efficiency of about 3%–4% due to incident photon-to-current
conversion efficiency of around 40%–45% at 500 nm. The low cell efficiency of ecda complexes
may be due to slow regeneration of the dye by electron donation from iodide following charge injec-
tion into TiO2.
New dyes of the type Ru(II)(bdmpp)(bpy) (where bdmpp is 2,6-bis(3,5-dimethyl-N-pyrazoyl)
pyridine, and bpy is 2,2′-bipyridine-4,4′-dicarboxylic acid) were synthesized by Falaras et al.
(2002). These compounds could be chemically anchored on TiO2 films via ester-like linkage involv-
ing carboxylato groups. These complexes were tested to act as potential molecular antenna in
dye-sensitized solar cells. The doctor blade technique was used to obtain opaque and transparent
nanocrystalline TiO2 thin-film electrodes, which were sensitized by these complexes and incorpo-
rated in a sandwich-type regenerative PEC solar cell containing 0.1 M LiI + 0.01 M I2 in propylene
carbonate. Platinized conductive glass was used as the counterelectrode. The overall energy conver-
sion efficiency was 1.72%.
Garcia et al. (2002) prepared a transparent photoanode by immobilizing cis-[(dcbH2)2Ru(CNpy)
(H2O)]2+ (dcbH2 is 4,4′-(CO2H)2-2,2′-bipyridine; CNpy is 4-cyanopyridine) in a TCO substrate
coated with nanocrystalline n-type TiO2 film. Charge recombination, quenching processes, time-
resolved experiments, and electron injection across the excited dye/semiconductor interface were
studied.
Vorobets et al. (2002) used concentrated polysulfide solutions as electrolytes in regenerative
PEC transducers. The equilibrium constants were determined, and the distribution of ion species in
the solutions was measured. The effect of electrolyte solution on the output characteristics of PEC
(based on cadmium selenide and cobalt sulfide) was also studied.
The pH dependence of sensitized photocurrent for porphyrin-derivatized planar TiO2 films
was studied by Watson et al. (2003). Porphyrin sensitizers like 5-(4-carboxyphenyl)-
10,15,20-trimesitylporphinatozinc(II), 5-(4-carboxyphenyl)-10,15,20-trimesitylporphine, and
5-(4-carboxyphenyl)-10,15,20-trimesitylporphinatoplatinum(II) were used. All three porphyrins
showed a 10-fold increase in the magnitude of sensitized photocurrent on acidification of the elec-
trolyte from pH 12 to pH 2.
Wang et al. (2003) employed silica nanoparticles to solidify ionic liquids. These ionic liquid–
based quasi-solid-state electrolytes were used in regenerative PEC cells. This yielded 7% efficiency
with an amphiphilic ruthenium polypyridyl photosensitizer. Jasieniak et al. (2004) compared the
photovoltaic performance of several porphyrin-derived TiO2 films in regenerative PEC cells with
the cells sensitized with a Ru(2,2′-bipyridyl-4,4′-dicarboxylate)2(NCS)2(N3) dye. They also studied
differences in efficiencies of the porphyrin light absorbers using porphyrin sensitizers tetrakis(3′,5′-
di-tert-butylphenyl)porphyrin, tetrakis(3′,5′-di-tert-butylphenyl)porphyrin zinc(II), and tetrakis(4′-
carboxyphenyl)porphyrin.
Stergiopoulos et al. (2005) incorporated new novel compounds in sandwich-type regenerative
PEC cells. A transition-metal complex with two terpy ligands [(2,2′:6′,2″-terpyridine-4′-iodophenyl)
(2,2′:6′,2″-terpyridine-4′-phenylphosphonic acid)-ruthenium(II)]dichloride, was used to sensitize
thin nanostructured SnO2 film electrodes. A high molecular mass poly(ethylene) oxide electrolyte
filled with titania and containing LiI and I2 was used to transport the current of the cell at the coun-
terelectrode. A continuous photocurrent (0.63 mA cm−2) and a photovoltage (290 mV) were pro-
duced by this cell under white light illumination. Incident photon-to-current conversion efficiencies
(IPCE) (16%) and energy conversion values (0.1%) were similar to those obtained with the standard
N3 dye under the same conditions.
Bergeron et al. (2005) implemented dye-sensitized mesoporous nanocrystalline SnO2 electrodes
and the pseudohalogen redox mediator (SeCN)2/SeCN− or (SCN)2/SCN− or the halogen redox media-
tor I3−/I−, for regenerative solar cells. The sensitizers used were Ru(deeb)(bpy)2(PF6)2, Ru(deeb)2(dpp)
(PF6)2, and Ru(deeb)2(bpz)(PF6)2, where deeb is 4,4′-diethylester, bpy is 2-2′-bipyridine, dpp is
Photoelectrochemical Cells 37

2,3-dipyridyl pyrazine, and bpz is bipyrazine. The donors present in the acetonitrile electrolyte
were used to reduce the oxidized sensitizers. A rate constant k > 108 s−1 was observed for sensitizer
regeneration with iodide as the donor. It was found that in regenerative solar cells, the IPCE and
open-circuit voltages were comparable for (SeCN)2/SeCN− and I3−/I− for all three sensitizers.
Bouroushian et al. (2006) synthesized polycrystalline (111) textured single CdSe and binary
CdSe/ZnSe thin films by electrodeposition. These films were used as active electrodes in regenera-
tive liquid-junction solar cells with aqueous sulfide–polysulfide or ferro–ferricyanide redox elec-
trolytes. The influence of ZnSe on the PEC properties of CdSe was studied. Corrosion effects and
stabilization of cells were also discussed.
Phenylenethynylene (PE) rigid linkers (para and meta) were used by Taratula et al. (2006) to
attach pyrene to the surface of TiO2 (anatase) and ZrO2 nanoparticle thin films through the two
COOH groups of an isophthalic acid unit. Photophysical properties of the compounds were influ-
enced by the length of the PE linkers and position of substitution (para or meta). The long wave-
length absorbance of the pyrene chromophore was shifted to the red with increasing conjugation,
and the extinction coefficient was also increased. Pyrene excimer acts as a sensitizer.
Brugnati et al. (2007) prepared and characterized a series of bipyridine and pyridyl-quinoline
Cu(I) complexes for their possible use as electron transfer mediators in regenerative PEC cells.
It was observed that the best-performing mediators produced maximum IPCEs of the order of
35%–40%.
Brennan et al. (2009) integrated a triethanolamine-protected silane, 1-(3′-amino) propylsilatrane,
with porphyrin- and ruthenium-based dyes and utilized it to link them to transparent semiconductor
nanoparticle metal oxide films. Silatrane reacts with the metal oxide to form strong, covalent silyl
ether bonds. Silatrane-functionalized dyes and analogous carboxylate-functionalized dyes were
used as visible light sensitizers for porous nanoparticle SnO2 photoanodes. The performances of the
dyes were compared in nonregenerative or regenerative PEC cells. NADH (β-nicotinamide adenine
dinucleotide) is used in nonregenerative cells as a sacrificial electron donor, and Hg2SO4/Hg is used
as a sacrificial cathode. In the regenerative cell, the iodide/triiodide redox couple was used. The
PEC cell efficiency was better improved with silatrane-based dyes than carboxylate-functionalized
dyes. Silatranes are more capable agents for bonding organic molecules to metal oxide surfaces.
Price and Maldonado (2009) prepared macroporous GaP photoelectrodes from nondegenerately
doped single-crystalline n-GaP (100). The PEC behaviors of planar and macroporous photoelec-
trodes were studied in nonaqueous regenerative PEC cells under potentiostatic control and using dry
acetonitrile containing ferrocene/ferrocenium. The changes in short-circuit photocurrents, open-
circuit photovoltages, and fill factors with increasing porosity resulted in improvement in the photo-
electrode efficiency of macroporous n-GaP.
Lee et al. (2009) developed a new procedure to prepare selenide (Se2−), which was used for
depositing CdSe quantum dots (QDs) over mesoporous TiO2 photoanodes by successive ionic layer
adsorption and a reaction (SILAR) process in ethanol. Optimization of QD-sensitized TiO2 films
was done by using a cobalt redox couple [Co(o-phen)3]2+/3+ in regenerative PEC cells. Over 4%
efficiency was achieved at 100 W m−2 with about 50% IPCE at its maximum on addition of a final
layer of CdTe. It was revealed that CdTe-terminated CdSe QD cells gave better charge collection
efficiencies compared to CdSe QD cells. They also prepared multilayered semiconductor (CdS/
CdSe/ZnS)–sensitized TiO2 mesoporous solar cells by the SILAR process (Lee et al. 2010). This
multicomponent sensitizer (CdS/CdSe/ZnS) was evaluated in a polysulfide electrolyte solution as a
redox mediator in regenerative PEC cells.
Onicha and Castellano (2010) focused on the effects of electrolyte composition, specifically
the role of Li+ and I− ions, on the resultant photovoltaic performance of dye-sensitized solar cells
based on a new Os(II) polypyridine complex, [Os(tBu3tpy)(dcbpyH2)(NCS)]PF6. Photophysical and
electrochemical characterization of this complex confirmed the suitability of the dye to serve as a
sensitizer for regenerative DSSCs on titania films. The photovoltaic performance of Os(II)-based
DSSCs could be enhanced by simply modifying the composition of redox electrolytes used in the
38 Solar Energy Conversion and Storage

operational sandwich cells. An abundance of I– played an important role in the effective regenera-
tion of oxidized surface-bound osmium sensitizers. The power conversion efficiency for an Os(II)-
based DSSC was calculated to be 4.7%.
Ruthenium(II) sensitizer (NBu4)[Ru(4,7-dpp)(dcbpyH)(NCS)2], (YS5) was synthesized, where
NBu4 is tetrabutylammonium; 4,7-dpp is 4,7-diphenyl-1,10-phenanthroline; and dcbpyH is the sin-
gly deprotonated surface anchoring derivative of 4,4′-dicarboxy-2,2′-bipyridine (dcbpyH2). This
was then incorporated into regenerative mesoscopic titania-based dye-sensitized solar cells (Sun
et al. 2010).
Xia et al. (2010) prepared, characterized, and anchored coordination compounds [Ru(NH3)5(eina)]
(PF6)2, [Ru(NH3)4(deeb)](PF6)2, and [Ru(en)2(deeb)](PF6)2, where eina is ethyl isonicotinate, deeb
is 4,4′-(CO2CH2CH3)2-2,2′-bipyridine, and en is ethylenediamine, to mesoporous nanocrystalline
(anatase) TiO2 thin films immersed in CH3CN at room temperature. The PEC performances of
[Ru(NH3)4(deeb)](PF6)2 and [Ru(en)2(deeb)](PF6)2 on TiO2 in regenerative solar cells were consistent
with excitation wavelength–dependent electron injection. Heuer et al. (2010) also prepared certain
Ru(II) compounds [Ru(bpy)2(mcbH)]2+ and [Ru(bpy)2(dafo)]2+, where bpy is 2,2′-bipyridine; mcbH
is 3-(CO2H)-2,2′-bipyridine; and dafo is 4,5-diazafluoren-9-one. These compounds were anchored
to nanocrystalline mesoporous TiO2 thin films for excited-state and interfacial electron transfer
studies. The IPCE was found to be lower for Ru(bpy)2(dafo)/TiO2.
Mesoporous SnO2 spheres were prepared for the first time by electrochemical anodization of tin
foil in basic media. Their structural elucidation indicated that these spherical particles consist of an
agglomeration of SnO2 nanocrystals, resulting in a high internal surface area, which would make
them a potential photoanode material for use in semiconductor-sensitized solar cells. After treating
SnO2 nanocrystals with aqueous TiCl4 solution, a thin layer of CdSe was coated by using ionic layer
adsorption and a reaction method. A power conversion efficiency of ∼1.91% was achieved in this
regenerative PEC cell after deposition of a ZnS passivation layer. Hagedorn et al. (2010) analyzed
the steady-state PEC responses of semiconductor nanowires in nonaqueous regenerative PEC cells.
The responses were used to determine the effect of width of the depletion region, relative to the
nanowire radius, on photogenerated carrier collection efficiency.
Johansson et al. (2011) synthesized, characterized, and contrasted three ruthenium compounds
(i.e., cis-Ru(dcbq)2(NCS)2, cis-Ru(dcbq)(bpy)(NCS)2, and cis-Ru(dcb)(bq)(NCS)2, where bpy is
2,2′-bipyridine, dcb is 4,4′-(CO2H)2-2,2′-bipyridine, bq is 2,2′-biquinoline, and dcbq is 4,4′-(CO2H)2-
2,2′-biquinoline) with the well-known N3 compound (i.e., cis-Ru(dcb)2(NCS)2) in dye-sensitized
solar cells. These compounds maintained the same cis-Ru(NCS)2 core with a variation in the energy
of the π* orbitals of the diimine ligand in the order bpy > dcb > bq > dcbq. The lowered π* orbit-
als resulted in enhanced red absorption as compared to N3. Sensitization from 400 to 900 nm was
realized with cis-Ru(dcb)(bq)(NCS)2 with HCl pretreated TiO2 in regenerative solar cells. Power
conversion efficiencies as high as 6.5% were obtained.
The diketonato-ruthenium(II)-polypyridyl sensitizers were anchored to nanocrystalline TiO2
films for light to electrical energy conversion in regenerative PEC cells by Islam et al. (2011).
Lewerenz (2011) described the principal design of nanoemitter solar cells and their applicability
in PEC solar cells that operate in the regenerative photovoltaic/photoelectrocatalytic mode as well
as in solid-state photovoltaics. Individual steps in the preparation of photovoltaic and photoelectro-
catalytic electrochemical solar cells with n- and p-type Si were discussed, and the electronic proper-
ties of nanoemitter solar cells at the solid–liquid phase boundary were also investigated. Klahr and
Hamann (2011) employed the atomic layer deposition technique to grow conformal thin films of
hematite on transparent conductive oxide substrates and used it as an electrode in regenerative PEC
cells. An increase in the photocurrent density and photovoltage was obtained by varying the pH and
redox potentials of the contacting electrolyte, which was attributed to increasing the built-in voltage.
Xiang et al. (2011) prepared p-type cuprous oxide photoelectrodes by the thermal oxidation of Cu
foils. These electrodes exhibited open-circuit voltages in excess of 800 mV in nonaqueous regenera-
tive PEC cells. Cuprous oxide gave an open-circuit voltage of 820 mV and a short-circuit current
Photoelectrochemical Cells 39

density of 3.1 mA cm−2 in contact with the decamethylcobaltocene+/0 (Me10CoCp2+/0) redox couple,


under simulated air mass 1.5 illumination. Han et al. (2011) explored a new class of thiocyanate-
free cyclometalated ruthenium sensitizers for sensitizing nanocrystalline TiO2 solar cells. A power
conversion efficiency of 4.76, a short-circuit photocurrent density of 11.21 mA cm−2, an open-circuit
voltage of 0.62 V, and a fill factor of 0.68 were obtained under standard air mass (AM) 1.5 sunlight.
In this cell, I−/I3− acetonitrile electrolyte was used. The complexes efficiently sensitized TiO2 over a
broad spectral range and showed an open-circuit potential of about 600 mV with a fill factor >0.70. Carli
et al. (2013) compared different poly(3,4-ethylenedioxythiophene (PEDOT)–based counterelectrodes,
obtained by potentiostatic electropolymerization of 3,4-ethylenedioxythiophene (EDOT) monomer on
fluorine tin oxide (FTO) surfaces, with the platinum- and gold-coated electrodes in order to evaluate
the potential use of PEDOT counterelectrodes in dye-sensitized PEC cells. A series of DSSC devices
utilizing Co(III)/(II) polypyridine redox mediators ([Co(bpy)3]3+/2+, [Co(phen)3]3+/2+, [Co(dtb)3]3+/2+,
where bpy is 2,2′-bipyridine, dtb is 4,4′di-tert-butyl-2,2′-bipyridine, and phen is 1,10-phenanthroline)
having distinct electrochemical characteristics were studied. Porous PEDOT/ClO4 counterelectrodes
have also been proven to possess sufficient electrocatalytic properties when paired with cobalt-based
redox mediators, making PEDOT-based counterelectrodes attractive for their use in DSC applications.
Mi et al. (2013) studied the behavior of WO3 photoanodes, in contact with a combination of four
anions (Cl−, CH3SO3−, HSO4−, and ClO4−) and three solvents (water, acetonitrile, and propylene
carbonate), to investigate the role of the semiconductor surface, the electrolyte, and redox kinetics
on the current density versus potential properties of n-type WO3. The internal quantum yield for
0.50 M tetra(n-butyl)ammonium perchlorate in propylene carbonate exceeded unity at excitation
wavelengths of 300–390  nm, which was an indication of current doubling. A regenerative PEC
cell based on the reversible redox couple B10Br10 −/2−  in acetonitrile, with a solution potential of
∼1.7 V versus the normal hydrogen electrode, showed an open-circuit photovoltage of 1.32 V under
100 mW cm−2.

4.2.2  Photoelectrolytic Solar Cells


The photoelectrosynthetic (photoelectrolytic or photocatalytic) cells utilize photon energy input (E ≥
Eg) to produce a net chemical change in the electrolyte solution (ΔG ≠ 0) when the reaction at the
counterelectrode is not exactly opposite of the hole transfer reaction at the illuminated semiconduc-
tor–liquid interface. The solar energy conversion efficiency for photoelectrolysis cells is

Energy stored as fuel − Electrical energy supplied


η= (4.3)
Incident solar energy
Two redox systems are present in photoelectrolytic cells. One redox system reacts with the holes at
the surface of the n-semiconductor electrode, while the other reacts with the electrons entering the
counterelectrode. Anodic and cathodic compartments need to be separated to prevent mixing of the
two redox couples. In case of n-SC, water is oxidized to oxygen at the semiconductor photoanode
and reduced to hydrogen at the cathode. The overall reaction is the cleavage of water by sunlight:

1
H 2O → H 2 + O 2 (4.4)
2

The commercial viability of dye-sensitized photoelectrosynthesis cells (DSPECs) depends on the


stability of the surface-bound molecular chromophores and catalysts. Water-stable, surface-bound
chromophores, catalysts, and assemblies are essential in DSPECs for the generation of solar fuels
by water splitting and CO2 reduction to CO, other oxygenates, or hydrocarbons.
Photopotentials were rarely above 0.6–0.8 V in PEC cells based on single junctions of semiconduc-
tors with solutions or metals. The higher photovoltages required multijunction cells involving multilayer
electrodes or a series connection of PEC cells. These principles were demonstrated with the instruments
40 Solar Energy Conversion and Storage

based on silicon spheres and Si p-n junctions contacting a solution via a noble metal layer. Reactions
considered include generation of Cl2 with reduction of O2 or generation of H2, the photobromination of
phenol, and the photochlorination of cyclohexene in acetonitrile (White et al. 1985).
Photoelectrosynthesis of dihydrogen through water splitting using S2− x (x = 1, 2, 3…) as an ano-
lyte, n-CdSe as the semiconductor electrode, and platinized-Pt as the cathode was reported by
Bhattacharyya et al. (1996). The semiconductor electrode was prepared by depositing a thin film of
CdSe by the RF sputtering technique on a stainless steel substrate. This is a first step in developing
a device for a rechargeable solar electrosynthetic cell.
Wadhawan et al. (2002) studied the mechanism of photoelectrochemically induced halex reac-
tions. They performed PEC reductions of para-bromonitrobenzene and 2,4-dibromonitrobenzene
in acetonitrile solutions. It was found that p-bromonitrobenzene followed a homogeneous ECrevCE
pathway in acetonitrile solutions containing tetrabutylammonium-based supporting electrolytes, but
when changing the supporting electrolyte to a salt of the tetramethylammonium cation, the mecha-
nism was changed qualitatively, and an ECEE pathway was observed. An ECECE mechanism was
observed on PEC reduction of 2,4-dibromonitrobenzene in acetonitrile solution containing support-
ing electrolytes derived from the tetrabutylammonium cation. When chloride-supporting electro-
lytes were used, then there was a light-induced rupture of a C–Br bond to a C–Cl bond. This halogen
change is a novel approach to halex reactions.
A strategy for the two-electron formation of C–C bonds with molecular catalysts anchored to
semiconductor nanocrystallites was demonstrated by Ardo et al. (2011). The semiconductor used
was the anatase polymorph of TiO2 present as a nanocrystalline, mesoporous thin film, and the
catalyst was cobalt meso-5,10,15,20-tetrakis(4-carboxyphenyl)porphyrin chloride, Co(TCPP)Cl.
Non-Nernstian two-electron transfer photocatalysis at iron protoporphyrin chloride-TiO2 interfaces
was performed in PESC. This non-Nernstian behavior was attributed to an environmentally depen-
dent potential drop across the molecule–semiconductor interface, and sustained photocurrents were
quantified in photoelectrosynthetic solar cells under forward bias.
Song et al. (2011) reported the use of [(Ru(bpy)2(4,4′-(PO3H2)2bpy)]2+ attached to TiO2 nanopar-
ticle films in a dye-sensitized photoelectrosynthesis cell for H2 production. Photoinduced electron
transfer in a chromophore-catalyst assembly anchored to TiO2 was investigated by Ashford et al.
(2012). A light-harvesting chromophore and a water oxidation catalyst were linked by a saturated
bridge designed to enable long-lived charge-separated states. Following excitation of the chromo-
phore, a rapid electron injection into TiO2 occurs, and an intra-assembly electron transfer occurs
on the subnanosecond time scale followed by microsecond–millisecond back-electron transfer from
the semiconductor to the oxidized catalyst. Song et al. (2012) also worked on the solar fuel forma-
tion in DSPECs.
The photostability of [RuII(bpy)2(4,4′-(PO3H2)2bpy)]Cl2(bpy is 4,4′-bipyridine) on nanocrystal-
line TiO2 and ZrO2 films was explored by Hanson et al. (2012). They examined stability by monitor-
ing visible light absorbance spectral changes, during 455 nm photolysis (475 mW cm−2) in a variety
of conditions relevant to dye-sensitized solar cells and dye-sensitized photoelectrosynthesis cells.
Enhancing the surface-binding stability of chromophores, catalysts, and chromophore-catalyst
assemblies attached to metal oxide surfaces is important for the development of photoelectrosyn-
thesis cells. Phosphonate-derivatized catalyst and molecular assembly provided a basis for water
oxidation on these surfaces in acidic solution but are unstable towards hydrolysis and loss from
surfaces as the pH was increased (Vannucci et al. 2013). This provided a hybrid approach to het-
erogeneous catalysis combining the advantages of systematic modifications possible by chemical
synthesis with heterogeneous reactivity. Alibabaei et al. (2013a) also described solar water splitting
in a dye-sensitized photoelectrosynthesis cell. A derivatized, core–shell nanostructured photoanode
was used with the core high surface area conductive metal oxide film coated with indium tin oxide
or antimony tin oxide with a thin outer shell of TiO2 formed by atomic layer deposition. A chromo-
phore-catalyst assembly, [(PO3H2)2bpy)2Ru(4-Mebpy-4-bimpy) Rub(tpy)(OH2)]4+, was attached to
the TiO2 shell, which combines both light absorber and water oxidation catalyst in a single molecule.
Photoelectrochemical Cells 41

Visible photolysis of the resulting core–shell assembly structure with a Pt cathode resulted in water
splitting into hydrogen and oxygen with an absorbed photon conversion efficiency of 4.4% at peak
photocurrent.
Norris et al. (2013) synthesized phosphonic acid derivatized bipyridine ligands and their ruthe-
nium complexes for DSPECs. This provided a stable chemical binding on metal oxide surfaces.
Hanson et al. (2013) used ALD of TiO2 on nanocrystalline TiO2 prefunctionalized with the dye
molecule [Ru(bpy)2(4,4′-(PO3H2)bpy)]2+(RuP), to stabilize surface-bound molecules. The result-
ing films were more photostable than untreated films, and the desorption rate constant decreased
exponentially with increased thickness of ALD TiO2 overlayers. The photodriven accumulation
of two oxidative equivalents at a single site was reported by cross electron transfer on TiO2 (Song
et al. 2013a). The TiO2 was coloaded with a ruthenium polypyridyl chromophore [Ru(bpy)2((4,4′-
(OH)2PO)2bpy)]2+ (RuIIP2+, bpy is 2,2′-bipyridine and ((OH)2PO)2-bpy is 2,2′-bipyridine-4,4′-
diyldiphosphonic acid) and a water oxidation catalyst [Ru(Mebimpy) ((4,4′-(OH)2PO-CH2)2bpy)
(OH2)]2+ (RuIIOH22+; Mebimpy is 2,6-bis(1-methylbenzimidazol-2-yl)pyridine; and (4,4′-(OH)2PO-
CH2)2bpy) is 4,4′-bis-methlylenephosphonato-2,2′-bipyridine). Steady-state illumination of coloaded
TiO2 photoanodes in a DSPEC configuration resulted in the makeup of -RuIIIP3+, -RuIIIOH2+, and
-RuIV-O2+, with -RuIV-O2+ formation preferred at high chromophore-to-catalyst ratios.
Thompson et al. (2013) investigated an integrated hybrid approach for making solar fuels in
DSPECs. Then, [Ru(bpy)3]2+* and its relatives provided a basis for exploring the energy gap law for
nonradiative decay, the role of molecular vibrations, and effects of solvent and medium on excited-
state properties. Alibabaei et al. (2013b) used metal oxide materials in dye-sensitized photoelectro-
synthesis cells to make solar fuels.
Song et al. (2013b) observed Li+ diffusion at the TiO2 interface in DSPEC. This cell worked in
aqueous solution at pH 4.5, and the rate constants for Li+ intercalation and release were observed
as 0.22  s–1 and 0.014  s–1, respectively. Both processes were considerably slower in the more vis-
cous solvent propylene carbonate. Release rate constants of Li+ were observed to be <2.0 × 10 −4 s−1.
Accumulation of Li+ under these conditions shifts conduction band/trap states to less negative
potentials, thus increasing electron lifetime in TiO2.
A sensitized Nb2O5 photoanode for hydrogen production in a dye-sensitized photoelectrosynthe-
sis cell was used by Luo et al. (2013). The conduction band potential is slightly positive (<0.1 eV)
in a T-phase orthorhombic Nb2O5 nanocrystalline film, relative to that of anatase TiO2. The H2
quantum yield and photostability measurements showed that Nb2O5 was comparable but not supe-
rior to TiO2 when ethylenediaminetetraacetate anion (EDTA4–) was added in DSPEC as a reductive
scavenger.
Dye-sensitized photoelectrosynthesis cells used in artificial photosynthesis require the assembly
of a chromophore and catalyst in close proximity on the surface of a transparent, high-bandgap
oxide semiconductor for integrated light absorption and catalysis. It was confirmed by controlled
potential electrolysis experiments that the surface-bound assemblies function as water oxidation
electrocatalysts. The electrochemical kinetics data indicated that the assemblies exhibit greater than
ten-fold rate enhancements as compared to the homogeneous catalyst alone (Ryan et al. 2014),
whereas Song et al. (2014) reported that light-driven dehydrogenation of benzyl alcohol to benz-
aldehyde and hydrogen occurred in a dye-sensitized photoelectrosynthesis cell. The photoanode
consists of nano-ITO, nano-ITO/TiO2, and mesoporous films of TiO2 nanoparticles in the DSPEC.
Excitations of chromophore and electron injection were followed by cross-surface electron transfer
activation of the catalyst to -RuIV=O2+, and in turn, benzyl alcohol was oxidized to benzaldehyde.
The injected electrons are transferred to a Pt electrode for the production of H2. Sustained absorbed
photon-to-current efficiency of 3.7% was achieved for benzyl alcohol dehydrogenation at the opti-
mized shell thickness, which amounts to an enhancement of ∼10 as compared to TiO2.
Ashford et al. (2014a) synthesized, characterized, and studied the electrochemical and photo-
physical properties of a series of Ru(II) polypyridyl complexes of the type [Ru(bpy)2(N-N)]2+ (bpy
is 2,2-bipyridine and N-N is a bidentate polypyridyl ligand). The nature of the N-N ligand was
42 Solar Energy Conversion and Storage

changed, either through increased conjugation or by incorporation of noncoordinating heteroatoms


in this series. Excited-state reduction potentials and Ru3+/2+ potentials were assessed in the context
of preparing low energy light absorbers for application in dye-sensitized photoelectrosynthesis cells.
Ashford et al. (2014b) reported water oxidation by an electropolymerized catalyst on derivatized
mesoporous metal oxide electrodes. This has certain applications in DSPECs. It was observed that
catalytic rate constants for water oxidation by the polymer films were similar to those for the phos-
phonated molecular catalyst on metal oxide electrodes, which indicates that the physical proper-
ties of the catalysts were not significantly changed in the polymer films. The results of controlled
potential electrolysis indicated sustained water oxidation over multiple hours with no decrease in
the catalytic current.
A helical peptide chromophore water oxidation catalyst assembly on a semiconductor surface
was loaded onto nanocrystalline TiO2 by Bettis et al. (2014). The oligoproline scaffold approach is
quite appealing due to its modular nature and helical tertiary structure. It maintained the controlled
relative positions of the chromophore and catalyst. Ultrafast transient absorption spectroscopy was
used to analyze the kinetics of the first photoactivation step for oxidation of water in the assembly.
This step in the water oxidation cycle of the chromophore-catalyst assembly anchored to TiO2 was
completed within 380 ps.

4.2.3  Photocatalytic Cells


In photoelectrocatalytic cells, the rate of reaction will increase, when ∆G < 0. Aqueous suspen-
sions composed of irradiated semiconductor particles may be considered to be an assemblage of
short-circuited microelectrochemical cells operating in the photocatalytic mode, for instance, the
photooxidation of organic compounds, or reactions with high activation energy.

N 2 + 3H 2 → 2NH 3 (4.5)

Photocatalytic oxidation of methyl red by TiO2 in a PEC cell with titania thin film loaded on tita-
nium as a photoanode has been studied by Shi et al. (2005). Degradation of methyl red was acceler-
ated in PEC cells due to a reduced recombination of photogenerated carriers by the separation of
the anodic and cathodic reactions. A thin SnO2 film was used as an interlayer between the substrate
and the TiO2 coating Ti/SnO2/TiO2 to obtain the assembled photoanodes, leading to an increased
separation efficiency of photogenerated carriers. The PEC water splitting for hydrogen production
using a combination of CIGS2 solar cell and RuO2 photocatalyst was studied by Dhere et al. (2004).
A PEC setup using a multiple bandgap combination of CuIn1–xGa xS2 (CIGS2) thin-film photovoltaic
cell and ruthenium oxide photocatalyst was presented.
The importance of bandgap energy and flat-band potential for application of modified TiO2 pho-
toanodes in water photolysis was reported by Radecka et al. (2008). The forbidden bandgap decides
the absorption spectrum, and the flat-band potential affects the recombination probability on water
photolysis. A three-electrode PEC cell with a TiO2 thin-film photoanode immersed in liquid electro-
lyte of variable pH was used. Titanium dioxide photoanodes doped with chromium (up to 16 at %)
and tin (up to 50 at %) were synthesized by RF reactive sputtering. The photoconversion efficiency
of TiO2 + 7.6 at % Cr was much smaller (ηc = 0.1%) than that of undoped TiO2 (ηc = 1.8%) and TiO2
doped with 8 at % of Sn (ηc = 1.0%).
Jeng et al. (2010) proposed a novel PEC cell for generation of hydrogen via photocatalytic water
splitting. This PEC cell is a membrane electrode assembly (MEA) integrated with Degussa P25
TiO2 powder as a model photocatalyst for the photoanode and Pt catalyst powder for the dark cath-
ode. It serves as an effective separator for the generated hydrogen and oxygen as well as a compact
photocatalytic reactor for water splitting. This novel PEC can be operated without the addition of
water in the cathode compartment, which showed improved photoconversion efficiency. Degussa
P25/BiVO4 mixed photocatalyst was found to significantly enhance the hydrogen generation.
Photoelectrochemical Cells 43

Preparation of TiO2 P-25 working electrodes on Ti substrates (TiO2/Ti) was reported by


Philippidis et al. (2010) by using pathogenic bacteria as a model in a novel batch PEC reactor.
The characterization and photoelectrocatalytic activity toward the inactivation of Escherichia coli
XL-1 blue (E. coli) colonies were studied. The flat-band potential (Vfb = –0.54 V) versus Ag/AgCl)
was determined by differential capacitance measurements. The photoelectrocatalytic inactivation of
E. coli colonies has been studied under artificial illumination in a novel photoelectrocatalytic reactor.
Sun et al. (2011) observed the effect of annealing temperature on the hydrogen production of
TiO2 nanotube arrays in a two-compartment PEC cell. Highly ordered TiO2 nanotube arrays with
4 μm length were prepared by a rapid anodization process in ethylene glycol electrolyte in order to
enhance the solar conversion efficiency. The highest photoconversion efficiency of 4.49% and maxi-
mum hydrogen production rate of 122 μmol/h·cm−2 has been observed using TiO2 NTs annealed
at 450°C.
Iodine-doped-poly(3,4-ethylenedioxythiophene)-modified Si nanowire one-dimensional (1D)
core–shell arrays were investigated by Yang et al. (2012) as efficient photocatalysts for solar
hydrogen. A new 1D core–shell strategy was developed for a hydrogen generation PEC cell. An
encouraging solar-to-chemical energy conversion efficiency was shown by this Si/iodine-doped
poly(3,4-ethylenedioxythiophene) (PEDOT) 1D nanocable array. Photocatalytic efficiency and sta-
bility of Si NW arrays could be increased by coating with iodine-doped PEDOT.
Rahman et al. (2012) studied the effects of doping (C or N) and co-doping (C + N) on the titania
coating for its application in solar water splitting. The bandgap (∼3.2 eV) of the TiO2 photocatalyst
is attributed to its relatively low photoactivity toward visible light. Therefore, efficient materials in
PEC cells should have a smaller bandgap (approximately 2.4 eV). The effect of dopant (C or N) and
co-dopant (C + N) was seen on the physical, structural, and photoactivity of TiO2 nano-thick coat-
ing. The type and amount of doping influenced the coating growth rate, structure, surface morphol-
ogy, and roughness. The photocurrent density (indirect indication of water-splitting performance)
of the C-doped photoanode was approximately 26% higher than undoped photoanode. The perfor-
mance of coating doped with N or N + C in the PEC cell was not found satisfactory due to their
higher charge recombination properties.
Preparation and characterization of co-doped TiO2 materials for solar light–induced current and
photocatalytic applications were done by Ganesh et al. (2012). A conventional coprecipitation and
sol-gel dip coating technique was used for the preparation of different amounts of co-doped TiO2
powders and thin films. The synthesized powders and thin films were subjected to thermal treat-
ments from 400°C to 800°C. The photocatalytic ability of the compounds was determined by the
degradation of methylene blue (MB). The results indicated that the co-doped TiO2 powder is mainly
of the anatase phase and possesses reasonably high specific surface area, low bandgap energy, and
flat-band potentials amenable to water oxidation in PEC cells. The 0.1 wt.% co-doped TiO2 compo-
sition provided the higher photocurrent, n-type semiconducting behavior, and higher photocatalytic
activity among various co-doped TiO2 compositions and pure TiO2.
Chao et al. (2012) studied visible light–driven photocatalytic and PEC properties of porous SnSx
(x = 1, 2) architectures. Porous SnS and SnS2 architectures on a large scale were synthesized using
a facile and template-free polyol refluxing process. It was observed that the as-synthesized SnS and
SnS2 products mainly consist of porous flower-like microstructures with reasonable BET surface
areas of 66  m2 g−1 and 33  m2 g−1, respectively. Both samples have excellent photosensitivity and
response with greatly enhanced Ion/off as high as 1.4 × 103. The potential applications of the SnSx
nanostructures in visible light–driven photocatalysis, high response photodetectors, and other opto-
electronic nanodevices have been investigated.
Generation of fuel from CO2-saturated liquids using a p-Si nanowire n-TiO2 nanotube array
PEC cell was observed by Latempa et al. (2012). Light-driven, electrically biased p-n junction PEC
cells immersed in an electrolyte of CO2-saturated 1.0 M NaHCO3 were investigated for generat-
ing hydrocarbon fuels. The PEC photocathode was composed of p-type Si nanowire arrays with
and without  copper sensitization, while the photoanode was composed of n-type TiO2 nanotube
44 Solar Energy Conversion and Storage

array films. The PEC cells convert CO2 into hydrocarbon fuels such as methane under bandgap
illumination, along with carbon monoxide and substantial rates of hydrogen generation due to water
photoelectrolysis. Methane and ethylene were formed at the combined rate of 201.5 nM/cm2-h at an
applied potential of –1.5 V versus Ag/AgCl with the addition of C3–C4 hydrocarbons. This tech-
nique provides a unique path for the photocatalytic reduction of CO2 with subsequent generation
of higher-order hydrocarbons and syngas constituents of carbon monoxide and hydrogen utilizing
Earth-abundant materials.
The water oxidation reactions based on photocatalyst suspension and PEC cells are currently
the most important field of research for sustainable energy. The effects of pH, crystallinity, and
grain size of tungsten trioxide (WO3) particles on photoconductivity were studied (Ho et al. 2012).
A simple hydrothermal route was employed to synthesize Cs-loaded WO3 (Cs-WO3) particles to
enhance the photochemical reactivity. A different photoreactivity performance in the PEC and
photocatalyst suspension system was shown by the photoanodes and photocatalyst based on the
Cs-WO3. This important fundamental insight can assist in the optimization of WO3 particles as the
photocatalyst and photoanode for future hybrid photocatalysis-electrolysis water-splitting systems.
Surface-modified WO3 particles were used for efficient water-splitting catalyst due to high photo-
catalytic capability.
Wongwanwattana et al. (2012) fabricated PEC cells based on metal-doped (Be and Fe) and
Pt-loaded nanostructured-TiO2 films as working electrodes for solar hydrogen production. Titanium
tetraisopropoxide was used as a precursor. The photocurrent density of Be-doped PEC without an
external applied potential has been observed as 0.32 mA cm−2 under illumination of 75 mW cm−2.
This device produced hydrogen by water photoelectrolysis at the rate of 0.1 mL h−1  cm−2, with a
photoconversion efficiency of 0.52%, whereas the maximum photocurrent density of Fe-doped PEC
was obtained as 0.80 mA cm−2 without an external applied potential and under the illumination of
100 mW cm−2 with photoconversion efficiency of 0.98%.
A facile polyol refluxing process has been used for the synthesis of tin sulfide (SnS), thick
10–20 nm nanoribbons and length up to several microns (Chao et al. 2013). The photoconductive
properties of the SnS nanoribbons were determined by collecting the samples into PEC cells, exhib-
iting excellent photosensitivity. The photocurrent density was 87 μA cm−2, which is the highest in
all the SnS photoelectrodes.
Gao et al. (2013) reported reduced graphene oxide–BiVO4 (GO-BiVO4) composite for PEC cells
and photocatalysis. The PEC cell working electrode was prepared by a doctor blade method on flu-
orine-doped tin oxide (FTO) coated glass. Graphene oxide–BiVO4 composites were synthesized by
hydrothermal reaction and, subsequently, reduced graphene oxide (RGO)-BiVO4 was obtained with
annealing in an N2 atmosphere. The RGO-BiVO4 films showed enhanced PEC properties under
visible light compared to pure BiVO4 film. It gave a high photocurrent response of 160 μA cm−2 and
quantum efficiency of over 1.81%. The enhanced PEC activity of RGO-BiVO4 could be explained
by its larger recombination resistance (Rrec) and longer electron lifetime (τ).
Jacobsson et al. (2013) reported that CIGS (CuIn xGa1−xSe2) can be utilized in the photocatalytic
reduction of water into hydrogen as an efficient absorber material. The efficiency was significantly
improved, and a photocurrent of 6 mA cm−2 was observed for the reduction reaction in the con-
figuration of a PEC cell by utilizing a solid-state p-n junction for charge separation and a catalyst
deposited on the surface. The separation between the charge carrier generation was demonstrated,
which takes place in the solar cell due to the catalysis. It takes place in the electrolyte leading to
improved stability, while keeping the essential functions of the processes. Photocurrents in excess
of 20 mA cm−2 were reached for the reduction half-reaction by incorporating appropriate charge
separation layers and optimizing the catalytic conditions at the surface of the electrodes.
The TiO2 branched nanorod arrays (TiO2 BNRs) were prepared (Su et al. 2013) with attached
plasmonic Au nanoparticles on the surface. It was observed that Au/TiO2 BNR composites exhibit
high photocatalytic activity in PEC water splitting. The unique structure of Au/TiO2 BNRs
shows enhanced activity with a photocurrent of 0.125 mA cm−2 under visible light (≥420 nm) and
Photoelectrochemical Cells 45

2.32 ± 0.1 mA cm−2 under AM 1.5 G illumination (100 mW cm−2). Furthermore, the Au/TiO2 BNRs


achieve the highest efficiency of ∼1.27% at a low bias of 0.50 V versus reversible hydrogen electrode
(RHE), indicating elevated charge separation and transportation efficiencies. It was also reported
that the high PEC performance was mainly due to the plasmonic effect of Au nanoparticles, which
enhances the visible light absorption, together with the large surface area, efficient charge separa-
tion, and high carrier mobility of the TiO2 BNRs. The carrier density of Au/TiO2 BNRs was nearly
six times higher than the pristine TiO2 BNRs as calculated by the Mott-Schottky plot.
Yang et al. (2013) suggested that dual cocatalysts are essential for developing photocatalytic
efficiency and PEC water-splitting reactions. It was observed that loading suitable dual cocatalysts
on semiconductors can significantly increase the photocatalytic activities of hydrogen and oxygen
evolution reactions, and it makes the overall water-splitting reaction possible.
Enhancing visible light PEC water splitting through transition-metal-doped TiO2 nanorod arrays
was reported by Wang et al. (2014). It was difficult to maintain high photocatalytic activity while
extending the photoresponse from the ultraviolet to the visible light region. Use of a transition-metal
doping treatment enhances the performance of TiO2 nanorods in the visible light region for PEC
water splitting. Then, Fe, Mn, and Co were used as dopants, and it was found that Fe doping is the
most effective among them. The photocurrent density of the Fe-TiO2 sample significantly increases
with bias voltage and reaches 2.92 mA cm−2 at 0.25 V versus Ag/AgCl, which is five times higher
than that of undoped TiO2. The Fe-TiO2 nanorod sample significantly improves the photoresponse,
not only in the UV region but also in the visible light region. The Fe-TiO2 nanorods have many
applications, such as solar water splitting, dye-sensitized solar cells, and photocatalysis.
The preparation of extremely smooth and boron-fluorine co-doped TiO2 nanotube arrays with
enhanced PEC and photocatalytic performance was reported by Li et al. (2014). Highly ordered
TiO2 nanotube arrays (BF-TNTs) were prepared in the unique NH4BF4-based electrolyte by the
anodization method. Boron and fluorine elements were simultaneously doped in the NH4BF4 elec-
trolyte to obtain BF-TNTs during the anodization process. Electrolyte NH4BF4 provides a smoother
effect on the tube walls of BF-TNTs. The BF-TNTs have better PEC properties and photocata-
lytic performance than do F-TNTs. Dual-template synthesis and PEC performance of hierarchical
porous zinc oxide were reported by Zhao et al. (2014). The ZnO was prepared by a sol-gel method.
A novel hierarchical porous ZnO was successfully synthesized through a sol-gel method. The pho-
toelectrochemical property of this hierarchical porous ZnO was investigated.
The unveiling of two electron transport modes in oxygen-deficient TiO2 nanowires and their
influence on PEC operation were investigated by Chen et al. (2014). One of the most promising
ways to enhance light-harvesting and photocatalytic efficiencies of PEC cells for water splitting is to
introduce oxygen vacancies (VO) into TiO2 materials, among others. The nature of electron transport
in VO-TiO2 nanostructures was not confirmed, especially in an operating device. A modulated pho-
tocurrent technique was used to study the electron transport property of VO-TiO2 nanowires (NWs).
The electron transport in pristine TiO2 displays a single trap-limited mode, whereas two electron
transport modes were found in VO-TiO2 NWs. The considerably higher diffusion coefficient of the
trap free transport mode grants a more rapid electron flow in VO-TiO2 NWs than that in pristine
TiO2 NWs.
ZnO was used as a photocatalytic electrode to TiO2 for solar-powered PEC electrolysis, and
H2 is generated by direct water splitting in such a cell with a metal cathode and a semiconducting
anode. The electrochemical potential and Fermi energy of the ZnO NR were calculated by the elec-
trochemical current density in acid and alkaline solutions via phenomenological thermodynamic
analysis. The ZnO NR has excellent potential for the storage of evolved H2, and it also appears to
operate as a hydrogen reservoir (Harinipriya et al. 2014).
Du et al. (2014) presented a novel biocathode-coupled PEC (Bio-PEC) integrating the advantages
of photocatalytic anode and biocathode. The electrochemical anodized TiO2 nanotube arrays fab-
ricated on Ti substrate were used as Bio-PEC anodes. The TiO2 nanotubes had inner diameters of
60–100 nm and wall thicknesses of about 5 nm. A pronounced photocurrent output (325 μA cm−2)
46 Solar Energy Conversion and Storage

under xenon illumination was reported as compared to dark conditions. A comparative study was
also carried out between the bio-PEC and PECs with Pt/C cathodes. The fill factors of bio-PEC and
brush-PEC (50 mg) were observed to be 39.87% and 43.06%, respectively. The charge-transfer resis-
tance of the biocathode was 13.10 ω, larger than the brush cathode with 50 mg Pt/C (10.68 ω), but
smaller than the brush cathode with 35 mg Pt/C (18.35 ω), indicating comparable catalytic activity
with the Pt/C catalyst. The biocathode was considered a promising alternative for the Pt/C catalyst
based on the performance and cost of the PEC system.
One metal oxide (titanium dioxide) and another silicon-based compound (silicon carbide) have
been given special attention as wide-bandgap semiconductor materials (Pessoa et al. 2015). Pessoa
et al. (2015) also presented material characteristics, synthesis methods, and recent photocatalytic
applications. The effects of the increase in efficiency of PEC devices that developed from the het-
erojunction of TiO2 and SiC were also observed.
Chen and Kamat (2014) used glutathione-capped metal (gold) nanoclusters as photosensitizers
for hydrogen generation in a PEC cell and a photocatalytic slurry reactor. The reversible reduction
(E 0  = −0.63 V vs. RHE) and oxidation (E 0  = 0.97 and 1.51 V vs. RHE) potentials of these metal
nanoclusters make them suitable for driving the water-splitting reaction. They also observed sig-
nificant photocurrent activity under visible light (400–500 nm) excitation, when a mesoscopic TiO2
film sensitized by Au x -GSH NCs was used as the photoanode with a Pt counterelectrode in aqueous
buffer solution (pH = 7). The rate of hydrogen production was 0.3 mol of hydrogen (h−1 g−1) of Au x -
GSH NCs, by sensitizing Pt/TiO2 nanoparticles with Au x -GSH NCs in an aqueous slurry system
under visible light. The rate of H2 evolution was significantly enhanced (∼5 times), when EDTA (as
a sacrificial donor) was introduced into the system.
A comparative study of photoelectrocatalytic and photocatalytic inactivation was made by Nie et al.
(2014). It was observed that the PEC inactivation was more effective to bacterial (E. coli K-12 and its
mutant E. coli BW25113) strains than the photocatalytic (PC) process. E. coli BW25113 showed higher
resistance than E. coli K-12 in both PEC and PC systems. It was found that h+ was the major reactive
species for PEC inactivation. Scanning electron microscopy images showed that the cells were severely
damaged, resulting in a leakage of the intracellular components during the PEC inactivation process.
The PEC inactivation efficiencies of both strains were enhanced in the presence of NaCl or NaBr.
Jin et al. (2014) prepared a flexible mesoporous TiO2 microspheres/cellulose acetate (TCA) hybrid
film as a recyclable photocatalyst with high performance, tunable size, and transparency. It was
obtained by a simple method of dispersing mesoporous TiO2 microspheres onto the surface of a free-
standing cellulose acetate (CA) film at room temperature. The PEC properties of the mesoporous
TiO2 microspheres were studied by configuring them as a simple self-powered PEC cell. It was also
observed that the TCA hybrid film displays excellent flexibility and favorable recyclable photocatalytic
activity for the decomposition of methylene blue (MB) solution under UV light irradiation. The pH
value of the solution has a more significant effect than temperature on the photoactivity of the sample.
Kim et al. (2014) reported the hierarchical In2O3:Sn/TiO2/CdS heterojunction nanowire array
photoanode (ITO/TiO2/CdS-nanowire array photoanode). It provides a short travel distance for the
charge carrier and a long light absorption pathway through the scattering effect. A comparison of
optical properties and device performance of the ITO/TiO2/CdS-nanowire array photoanode with
the TiO2 nanoparticle/CdS photoanode has been presented. The photocatalytic properties for water
splitting were also observed in the presence of a sacrificial agent such as SO32− and S2− ions. The
ITO/TiO2/CdS-nanowire array photoanode exhibits a photocurrent density of 8.36 mA cm−2, under
illumination (AM 1.5 G, 100 mW cm−2), at 0 V versus Ag/AgCl, which is four times higher than
that of the TiO2 nanoparticle/CdS photoanode. It was concluded that improved light-harvesting and
charge collection properties, due to the increased light absorption pathway and reduced electron
travel distance by ITO nanowire, lead to enhancement of the PEC performance.
Guo et al. (2014) prepared ZnO/Cu2S core–shell nanorods from ZnO NRs for PEC water splitting
through a versatile hydrothermal chemical conversion method (H-ZnO/Cu2S core–shell NRs) and suc-
cessive ionic layer adsorption and reaction method (S-ZnO/Cu2S core–shell NRs). The photoelectrode
Photoelectrochemical Cells 47

was composed of a core–shell structure, where the core portion is ZnO NRs and the shell portion
is Cu2S nanoparticles sequentially located on the surface. It was observed that the ZnO NRs array
provides a fast electron transport pathway due to its high electron mobility properties. This PEC sys-
tem produced very high photocurrent density and photoconversion efficiency under 1.5 AM irradia-
tion for hydrogen generation. It was demonstrated that H-ZnO/Cu2S core–shell NRs exhibit a much
higher photocatalytic activity than S-ZnO/Cu2S core–shell NRs. The photocurrent density and pho-
toconversion efficiency of H-ZnO/Cu2S core–shell NRs were up to 20.12 mA cm−2 at 0.85 V versus
SCE and 12.81% at 0.40 V versus SCE, respectively.
Li et al. (2015) constructed ternary CdS/reduced graphene oxide/TiO2 nanotube array hybrids
for enhanced visible light–driven PEC and photocatalytic activity. The coupling technique of elec-
trophoretic deposition (EPD) was used for the synthesis of ternary nanocomposite photoelectrodes
composed of CdS nanocrystallites, reduced graphene oxide (RGO), and TiO2 nanotube arrays
(TNTs). The ternary CdS/RGO/TNTs hybrids show more visible light–driven PEC and photocata-
lytic activity because the outer layer of CdS acts as a sensitizer for trapping substantial photons
from the visible light. The middle layer of RGO not only serves as a transporter for suppressing the
recombination of photogenerated carriers, but also acts as a green sensitizer for increasing visible
light absorption. The inner TNTs with narrowed bandgap collect the hot electrons form the visible
light absorption, while CdS and RGO participate in subsequent redox reactions for hydrogen pro-
duction and degradation of organic pollutants.
The effect of NH4F concentration and controlled charge consumption on the photocatalytic
hydrogen generation on TiO2 nanotube arrays was studied by Yujing et al. (2015). Electrochemical
anodization in ethylene glycol–based electrolytes with various NH4F concentrations was used for
fabricating self-organized TiO2 nanotube arrays (TiO2 NTs) for hydrogen evolution. Hydrogen pro-
duction by photocatalytic water splitting was performed in a two-compartment PEC cell, and in
this process, no applied voltage was used. The impacts of NH4F concentration on the morphologi-
cal structure, PEC performance, and hydrogen evolution of TiO2 samples were analyzed. The TiO2
anodized with 0.50 wt% of NH4F concentration for 60 min exhibited the highest hydrogen evolution
of 2.53 mL h−1  cm−2 and maximum photoconversion efficiency of 4.39%. Another series of TiO2
nanotube arrays samples with the equal charge consumption (designated as TiO2 NTs-EC) was pre-
pared. It was observed that hydrogen production and PEC properties of TiO2 NTs-EC samples have
increased. It was concluded that anodization charge density plays an important role in hydrogen
generation by TiO2 NTs.

4.3  PEC CELLS AS STORAGE CELLS


The PEC solar cells have the potential not only to convert but also store incident solar energy.
Efficient photochemical storage and conversion require better functional performance of the sepa-
rate cell components and system compatibility.
Some parameters should be optimized for obtaining maximum efficiency. These are

• Stability of the photoelectrode


• Stability of the counterelectrode
• Stability of the electrolyte
• Reversibility
• High photoelectrode-conversion efficiency
• Minimization of loss in light intensity reaching the photoelectrode
• High current and potential efficiency of the redox storage process
• High energy capacity of the redox storage

The PEC cells not only generate electricity, but they also split water efficiently into hydrogen and
oxygen. Water splitting is a welcome step to produce hydrogen, which is being advocated as the
48 Solar Energy Conversion and Storage

fuel of the future. A number of PEC cells have been reported with varied materials in the form of
either electrodes or electrolytes. Various efforts have been made to achieve efficient and stable PEC
cells, including surface modification of the photoelectrode, electrolyte modification, photoetching
of layered semiconductors, new configurations of PEC solar cells, dye sensitization, and so on. Still
there are many opportunities available to develop newer materials for the future.

REFERENCES
Alebbi, M., C. A. Bignozzi, T. A. Heimer, G. M. Hasselmann, and G. J. Meyer. 1998. The limiting role of iodide
oxidation in cis-Os(dcb)2(CN)2/TiO2 photoelectrochemical cells. J. Phys. Chem. B. 102: 7577–7581.
Alibabaei, L., H. Luo, R. L. House, P. G. Hoertz, R. Lopez, and T. J. Meyer. 2013a. Applications of metal oxide
materials in dye sensitized photoelectrosynthesis cells for making solar fuels: Let the molecules do the
work. J. Mater. Chem. A. 1: 4133–4145.
Alibabaei, L., M. K. Brennaman, M. R. Norris, B. Kalanyan, W. Song, M. D. Losego, et al. 2013b. Solar water
splitting in a molecular photoelectrochemical cell. Proc. Nat. Acad. Sci. USA. 110: 20008–20013.
Ang, P. G. P., A. F. Sammells, Y. Sun, A. C. Onicha, M. Myahkostupov, and F. N. Castellano. 1983. A one-volt
p-InP/n-CdSe regenerative photoelectrochemical cell. J. Electrochem. Soc. 130: 1784–1786.
Ardo, S., D. Achey, A. J. Morris, M. Abrahamsson, and G. J. Meyer. 2011. Non-Nernstian two-electron trans-
fer photocatalysis at metalloporphyrin–TiO2 interfaces. J. Am. Chem. Soc. 133: 16572–16580.
Argazzi, R., C. A. Bignozzi, G. M. Hasselmann, and G. J. Meyer. 1998. Efficient light-to-electrical energy
conversion with dithiocarbamate-ruthenium polypyridyl sensitizers. Inorg. Chem. 37: 4533–4537.
Ashford, D. L., C. R. K. Glasson, M. R. Norris, J. J. Concepcion, S. Keinan, M. K. Brennaman, et al. 2014a.
Controlling ground and excited state properties through ligand changes in ruthenium polypyridyl com-
plexes. Inorg. Chem. 53: 5637–5646.
Ashford, D. L., A. M. Lapides, A. K. Vannucci, K. Hanson, D. A. Torelli, D. P. Harrison, et al. 2014b. Water
oxidation by an electropolymerized catalyst on derivatized mesoporous metal oxide electrodes. J. Am.
Chem. Soc. 136: 6578–6581.
Ashford, D. L., W. Song, J. J. Concepcion, C. R. K. Glasson, M. K. Brennaman, M. R. Norris, et al. 2012.
Photoinduced electron transfer in a chromophore-catalyst assembly anchored to TiO2. J. Am Chem Soc.
134: 19189–19198.
Bergeron, B. V., A. Marton, G. Oskam, and G. J. Meyer. 2005. Dye-sensitized SnO2 electrodes with iodide and
pseudohalide redox mediators. J. Phys. Chem. B. 109: 937–943.
Bettis, S. E., D. M. Ryan, M. K. Gish, L. Alibabaei, T. J. Meyer, M. L. Waters, et al. 2014. Photophysical
characterization of a helical peptide chromophore-water oxidation catalyst assembly on a semiconductor
surface using ultrafast spectroscopy. J. Phys. Chem. C. 118: 6029–6037.
Bhattacharyya, R. G., D. P. Mandal, S. C. Bera, and K. K. Rohatgi-Mukherjee. 1996. Photoelectrosynthesis
of dihydrogen via water-splitting using S2−x (x = 1, 2, 3...) as an anolyte: A first step for a viable solar
rechargeable battery. Int. J. Hydrogen Energy. 21: 343–347.
Bhavani, N. K., D. Vijayalakshmi, and S. Seshan. 1986. Photochemical conversion and storage of solar energy.
Energy Manage. 10: 27–32.
Bicelli, L. P., G. Razzini, N. Romeo, and V. Canevari. 1985. Photoelectrochemical characterization of
n-CuInSe2 films prepared by quasi-rheotaxy. Surface Technol. 25: 327–334.
Bouroushian, M., D. Karoussos, and T. Kosanovic. 2006. Photoelectrochemical properties of electrodeposited
CdSe and CdSe/ZnSe thin films in sulphide-polysulphide and ferro-ferricyanide redox systems. Solid
State Ionics. 177: 1855–1859.
Brennan, B. J., A. E. Keirstead, P. A. Liddell, S. A. Vail, T. A. Moore, A. L. Moore, et al. 2009. 1-(3′-amino)
propylsilatrane derivatives as covalent surface linkers to nanoparticulate metal oxide films for use in
photoelectrochemical cells. Nanotechnology. 20: 505203.
Brugnati, M., S. Caramori, S. Cazzanti, L. Marchini, R. Argazzi, and C. A. Bignozzi. 2007. Electron transfer
mediators for photoelectrochemical cells based on Cu(I) metal complexes. Int. J. Photoenergy. 2007: 80756.
Carli, S., E. Busatto, S. Caramori, R. Boaretto, R. Argazzi, C. J. Timpson, et al. 2013. Comparative evaluation
of catalytic counter electrodes for Co(III)/(II) electron shuttles in regenerative photoelectrochemical
cells. J. Phys. Chem. C. 117: 5142–5153.
Carlsson, P., and B. Holmström. 1986. Photoelectrochemical cells: Laboratory determination of solar conver-
sion efficiencies. Solar Energy. 36: 151–157.
Cattarin, S., G. Zotti, M. M. Musiani, and G. Mengoli. 1986. Photoelectrochemistry of para-diethynylbenzene
copper(I) acetylides deposited as thin films onto copper electrodes. J. Electroanal. Chem. 207: 247–261.
Photoelectrochemical Cells 49

Chao, J., Z. Wang, X. Xu, Q. Xiang, W. Song, G. Chen, et al. 2013. Tin sulfide nanoribbons as high perfor-
mance photoelectrochemical cells, flexible photodetectors and visible-light-driven photocatalysts. RSC
Adv. 3: 2746–2753.
Chao, J., Z. Xie, X. Duan, Y. Dong, Z. Wang, J. Xu, et al. 2012. Visible-light-driven photocatalytic and pho-
toelectrochemical properties of porous SnSx (x = 1,2) architectures. Cryst. Eng. Comm. 14: 3163–3168.
Chen, H., Z. Wei, K. Yan, Y. Bai, and S. Yang. 2014. Unveiling two electron-transport modes in oxygen-
deficient TiO2 nanowires and their influence on photoelectrochemical operation. J. Phys. Chem. Lett.
5: 2890–2896.
Chen, Y.-S., and P. V. Kamat. 2014. Glutathione-capped gold nanoclusters as photosensitizers. Visible light-
induced hydrogen generation in neutral water. J. Am. Chem. Soc. 136: 6075–6082.
Dhere, N. G., A. H. Jahagirdar, U. S. Avachat, and A. A. Kadam. 2004. Photoelectrochemical water splitting
for hydrogen production using combination of CIGS2 solar cell and RuO2 photocatalyst. Thin Solid
Films. 481–482: 462–465.
Du, Y., Y. Feng, Y. Qu, J. Liu, N. Ren, and H. Liu. 2014. Electricity generation and pollutant degradation using
a novel biocathode coupled photoelectrochemical cell. Environ. Sci. Technol. 48: 7634–7641.
Falaras, P., K. Chryssou, T. Stergiopoulos, I. Arabatzis, G. Katsaros, V. J. Catalano, et al. 2002. Dye-
sensitization of titanium dioxide thin films by Ru(II)-bpp-bpy complexes. Proc. SPIE–Int. Soc. Optical
Eng. 4801: 125–135.
Falaras, P., A. P. Xagas, and A. H.-L. Goff. 1998. Synthesis and characterization of dichloro(2,2′-bipyridyl-
4,4′dicarboxylate)bis(triphenylphosphine) ruthenium(II) for efficient photosensitization of titanium
oxide. New J. Chem. 22: 557–558.
Forouzan, F., and S. Licht. 1995. Solution-modified n-GaAs/aqueous polyselenide photoelectrochemistry.
J. Electrochem. Soc. 142: 1539–1545.
Fujishima, and K. Honda. 1972. Electrochemical photolysis of water at a semiconductor electrode. Nature.
238: 37–38.
Ganesh, I., A. K. Gupta, P. P. Kumar, P. S. Chandra Sekhar, K. Radha, G. Padmanabham, et al. 2012.
Preparation and characterization of Co-doped TiO2 materials for solar light induced current and photo-
catalytic applications. Mater. Chem. Phys. 135: 220–234.
Gao, L., F. Qu, and X. Wu. 2013. Reduced graphene oxide-BiVO4 composite for enhanced photoelectrochemi-
cal cell and photocatalysis. Sci. Adv. Mater. 5: 1485–1492.
Garcia, C. G., and N. Y. M. Iha. 2001. Photoelectrochemical solar cells using [(dcbH2)2 RuLL′], L,L′ = substi-
tuted pyridines, as nanocrystalline TiO2 sensitizers. Int. J. Photoenergy. 3: 130–135.
Garcia, C. G., N. Y. M. Iha, R. Argazzi, and C. A. Bignozzi. 1998. 4-Phenylpyridine as ancillary ligand
in ruthenium(II) polypyridyl complexes for sensitization of n-type TiO2 electrodes. J. Photochem.
Photobiol. A: Chem. 115: 239–242.
Garcia, C. G., A. K. Nakano, C. J. Kleverlaan, and N. Y. M. Iha. 2002. Electron injection versus charge recom-
bination in photoelectrochemical solar cells using cis-[(dcbH2)2Ru(CNpy)(H2O]Cl2 as a nanocrystalline
TiO2 sensitizer. J. Photochem. Photobiol. A: Chem. 151: 165–170.
Guay, D., R. Cote, R. Marques, J. P. Dodelet, M. F. Lawrence, D. Gravel, et al. 1987. High short-circuit pho-
tocurrents obtained with a phthalocyanine in a regenerative photoelectrochemical cell. J. Electrochem.
Soc. 134: 2942–2943.
Guo, K., X. Chen, J. Han, and Z. Liu. 2014. Synthesis of ZnO/Cu2S core/shell nanorods and their enhanced
photoelectric performance. J. Sol-Gel Sci. Tech. 72: 92–99.
Hagedorn, K., C. Forgacs, S. Collins, and S. Maldonado. 2010. Design considerations for nanowire hetero-
junctions in solar energy conversion/storage applications. J. Phys. Chem. C. 114: 12010–12017.
Han, L., S. P. Singh, A. Islam, and M. Yanagida. 2011. Development of a new class of thiocyanate-free cyclo-
metalated ruthenium(II) complex for sensitizing nanocrystalline TiO2 solar cells. Int. J. Photoenergy.
2011: 520848.
Hanson, K., M. K. Brennaman, H. Luo, C. R. K. Glasson, J. J. Concepcion, W. Song, et al. 2012. Photostability
of phosphonate-derivatized, RuII polypyridyl complexes on metal oxide surfaces. ACS Appl. Mater.
Interfaces. 4: 1462–1469.
Hanson, K., M. D. Losego, B. Kalanyan, G. N. Parsons, and T. J. Meyer. 2013. Stabilizing small molecules on
metal oxide surfaces using atomic layer deposition. Nano Lett. 13: 4802–4809.
Harinipriya, S., B. Usmani, D. J. Rogers, V. E. Sandana, F. H. Teherani, Lusson, et al. 2014. ZnO nanorod
electrodes for hydrogen evolution and storage (Conference Paper). J. Sol-Gel Sci. Technol. 72:
92–99.
Heimer, T. A., S. T. D’Arcangelis, F. Farzad, J. M. Stipkala, and G. J. Meyer. 1996. An acetylacetonate-based
semiconductor-sensitizer linkage. Inorg. Chem. 35: 5319–5324.
50 Solar Energy Conversion and Storage

Heuer, W. B., H.-L. Xia, M. Abrahamsson, Z. Zhou, S. Ardo, A. A. N. Sarjeant, et al. 2010. Reaction of RuII
diazafluorenone compound with nanocrystalline TiO2 thin film. Inorg. Chem. 49: 7726–7734.
Ho, G. W., K. J. Chua, and D. R. Siow. 2012. Metal loaded WO3 particles for comparative studies of photoca-
talysis and electrolysis solar hydrogen production. Chem. Eng. J. 181–182: 661–666.
Islam, A., S. P. Singh, and L. Han. 2011. Synthesis and application of new ruthenium complexes contain-
ing β-diketonato ligands as sensitizers for nanocrystalline TiO2 solar cells. Int. J. Photoenergy. 2011:
204639.
Islam, A., H. Sugihara, K. Hara, L. P. Singh, R. Katoh, M. Yanagida, et al. 2001a. Dye sensitization of nano-
crystalline titanium dioxide with square planar platinum(II) diimine dithiolate complexes. Inorg. Chem.
40: 5371–5380.
Islam, A., H. Sugihara, K. Hara, L. P. Singh, R. Katoh, M. Yanagida, et al. 2001b. Sensitization of nanocrys-
talline TiO2 film by ruthenium(II) diimine dithiolate complexes. J. Photochem. Photobiol. A: Chem.
145: 135–141.
Jacobsson, T. J., C. Platzer-Björkman, M. Edoff, and T. Edvinsson. 2013. CuInxGa1-xSe2 as an efficient photo-
cathode for solar hydrogen generation. Int. J. Hydrogen Energy. 38: 15027–15035.
Jasieniak, J., M. Johnston, and E. R. Waclawik. 2004. Characterization of a porphyrin-containing dye-sensi-
tized solar cell. J. Phys. Chem. B. 108: 12962–12971.
Jeng, Ki.-T., Y.-C. Liu, Y.-F. Leu, Y.-Z. Zeng, J.-C. Chung, and T.-Y. Wei. 2010. Membrane electrode assembly-
based photoelectrochemical cell for hydrogen generation. Int. J. Hydrogen Energy. 35: 10890–10897.
Jin, X., J. Xu, X. Wang, Z. Xie, Z. Liu, B. Liang, et al. 2014. Flexible TiO2/cellulose acetate hybrid film as a
recyclable photocatalyst. RSC Adv. 4: 12640–12648.
Johansson, P. G., J. G. Rowley, A. Taheri, G. J. Meyer, S. P. Singh, A. Islam, et al. 2011. Long-wavelength
sensitization of TiO2 by ruthenium diimine compounds with low-lying π* orbitals. Langmuir. 27:
14522–14531.
Keita, B., and L. Nadjo. 1984. Electrochemistry and photoelectrochemistry of sodium 9,10-anthraquinone-2,6.
disulfonate in aqueous media. Application to rechargeable solar cells and to the synthesis of hydrogen
peroxide. J. Electroanal. Chem. 163: 171–188.
Kelly, C. A., F. Farzad, D. W. Thompson, J. M. Stipkala, and G. J. Meyer. 1999. Cation-controlled interfacial
charge injection in sensitized nanocrystalline TiO2. Langmuir. 15: 7047–7054.
Kim, J. S., H. S. Han, S. Shin, G. S. Han, H. S. Jung, K. S. Hong, et al. 2014. In2O3:Sn/TiO2/CdS heterojunction
nanowire array photoanode in photoelectrochemical cells. Int. J. Hydrogen Energy. 39: 17473–17480.
Klahr, B. M., and T. W. Hamann. 2011. Current and voltage limiting processes in thin film hematite elec-
trodes. J. Phys. Chem. C. 115: 8393–8399.
Kline, G., K. Kam, D. Canfield, and B. A. Parkinson. 1981. Efficient and stable photoelectrochemical cells
constructed with WSe2 and MoSe2 photoanodes. Solar Energy Mater. 4: 301–308.
Kline, G., K. K. Kam, R. Ziegler, and B. A. Parkinson. 1982. Further studies of the photoelectrochemical
properties of the group VI transition metal dichalcogenides. Solar Energy Mater. 6: 337–350.
Kohl, P. A., and A. J. Bard. 1979. Semiconductor electrodes-liquid junction photovoltaic cells based on n-GaAs
electrodes and acetonitrile solutions. J. Electrochem. Soc. 126: 603–608.
Ladouceur, M., J. P. Dodelet, G. Tourillon, L. Parent, and S. Dallaire. 1990. Plasma-sprayed semiconduc-
tor electrodes: Photoelectrochemical characterization and NH3 photoproduction by substoichiometric
tungsten oxides. J. Phys. Chem. 94: 4579–4587.
Latempa, T. J., S. Rani, N. Bao, and C. A. Grimes. 2012. Generation of fuel from CO2 saturated liquids using
a p-Si nanowire n-TiO2 nanotube array photoelectrochemical cell. Nanoscale. 4: 2245–2250.
Lee, H. J., J. Bang, J. Park, S. Kim, and S.-M. Park. 2010. Multilayered semiconductor (CdS/CdSe/ZnS)-
sensitized TiO2 mesoporous solar cells: All prepared by successive ionic layer adsorption and reaction
processes. Chem. Mater. 22: 5636–5643.
Lee, H., M. Wang, P. Chen, D. R. Gamelin, S. M. Zakeeruddin, M. Grätzel, et al. 2009. Efficient CdSe quan-
tum dot-sensitized solar cells prepared by an improved successive ionic layer adsorption and reaction
process. Nano Lett. 9: 4221–4227.
Lewerenz, H. J. 2011. Operational principles of electrochemical nanoemitter solar cells for photovoltaic and
photoelectrocatalytic applications. J. Electroanal. Chem. 662: 184–195.
Li, H., Z. Xia, J. Chen, L. Lei, and J. Xing. 2015. Constructing ternary CdS/reduced graphene oxide/TiO2
nanotube arrays hybrids for enhanced visible-light-driven photoelectrochemical and photocatalytic
activity. J. Appl. Catal. B: Environ. 168–169: 105–113.
Li, H., J. Xing, Z. Xia, and J. Chen. 2014. Preparation of extremely smooth and boron-fluorine co-doped TiO2
nanotube arrays with enhanced photoelectrochemical and photocatalytic performance. Electrochim.
Acta. 139: 331–336.
Photoelectrochemical Cells 51

Licht, S. 2001. Multiple band gap semiconductor/electrolyte solar energy conversion. J. Phys. Chem. B. 105:
6281–6294.
Licht, S., O. Khaselev, P. A. Ramakrishnan, D. Faiman, E. A. Katz, A. Shames, et al. 1998a. Fullerene photo-
electrochemical solar cells. Sol. Energy Mater. Solar Cells. 51: 9–19.
Licht, S., O. Khaselev, P. A. Ramakrishnan, T. Soga, and M. Umeno. 1998b. Multiple bandgap photoelectrochem-
istry: Bipolar semiconductor ohmic regenerative electrochemistry. J. Phys. Chem. B. 102: 2536–2545.
Licht, S., O. Khaselev, P. A. Ramakrishnan, T. Soga, and M. Umeno. 1998c. Multiple bandgap photoelec-
trochemistry: Inverted semiconductor Ohmic regenerative electrochemistry. J. Phys. Chem. B. 102:
2546–2554.
Licht, S., O. Khaselev, T. Soga, and M. Umeno. 1998d. Multiple bandgap photoelectrochemistry: Energetic
configurations for solar energy conversion. Electrochem. Solid State Lett. 1: 20–23.
Licht, S., P. A. Ramakrishnan, D. Faiman, E. A. Katz, A. Shames, and S. Goren. 1998e. Photoaction, tem-
perature and O2 depletion effects in fullerene photoelectrochemical solar cells. Sol. Energy Mater. Solar
Cells. 56: 45–55.
Luo, H., W. Song, P. G. Hoertz, K. Hanson, R. Ghosh, S. Rangan, et al. 2013. A sensitized Nb2O5 photoanode
for hydrogen production in a dye-sensitized photoelectrosynthesis cell. Chem. Mater. 25: 122–131.
Matthews, D., P. Infelta, and M. Grätzel. 1996. Calculation of the photocurrent-potential characteristic for
regenerative, sensitized semiconductor electrodes. Sol. Energy Mater. Solar Cells. 44: 119–155.
McEvoy, A. J. 2005. Photoelectrochemical solar cells. In T. Markvart and L. Castaner (Eds), Solar Cells:
Materials, Manufacture and Operation. Oxford: Elsevier, pp. 395–417.
Memming, R. 1978. Energy production by photoelectrochemical processes. Philip. Tech. Rev. 38: 160–177.
Mi, Q., R. H. Coridan, B. S. Brunschwig, H. B. Gray, and N. S. Lewis. 2013. Photoelectrochemical oxidation
of anions by WO3 in aqueous and nonaqueous electrolytes. Energy Environ. Sci. 6: 2646–2653.
Miller, B. 1984. Charge transfer and corrosion processes at III-V semiconductor/electrolyte interfaces.
J. Electroanal. Chem. 168: 91–100.
Neumann-Spallart M., and K. Kalyanasundaram. 1981. Photoelectrochemical cells with polycrystalline cad-
mium sulfide as photoanodes. Ber. Bunsen. Phys. Chem. 85: 1112–1117.
Nie, X., G. Li, M. Gao, H. Sun, X. Liu, H. Zhao, et al. 2014. Comparative study on the photoelectrocatalytic
inactivation of Escherichia coli K-12 and its mutant Escherichia coli BW25113 using TiO2 nanotubes as
a photoanode. Appl. Catal. B: Environ. 147: 562–570.
Norris, M. R., J. J. Concepcion, C. R. K. Glasson, Z. Fang, A. M. Lapides, D. L. Ashford, et al. 2013. Synthesis
of phosphonic acid derivatized bipyridine ligands and their ruthenium complexes. Inorg. Chem. 52:
12492–12501.
Onicha, A. C., and F. N. Castellano. 2010. Electrolyte-dependent photovoltaic responses in dye-sensitized
solar cells based on an osmium(II) dye of mixed denticity. J. Phys. Chem. C. 114: 6831–6840.
Osaka, T., N. Hirota, T. Hayashi, and S. S. Eskildsen. 1985. Characteristics of photoelectrochemical cells with
iron oxide/n-Si heterojunction photoanodes. Electrochim. Acta. 30: 1209–1212.
Papageorgiou, N., Y. Athanassov, M. Armand, P. Bonhôte, H. Pettersson, A. Azam, et al. 1996. The perfor-
mance and stability of ambient temperature molten salts for solar cell applications. J. Electrochem. Soc.
143: 3099–3108.
Park, J. K., H. R. Lee, J. Chen, H. Shinokubo, A. Osuka, and D. Kim. 2008. Photoelectrochemical properties
of doubly β-functionalized porphyrin sensitizers for dye-sensitized nanocrystalline-TiO2 solar cells.
J. Phys. Chem. C. 112: 16691–16699.
Pessoa, R. S., M. A. Fraga, L. V. Santos, M. Massi, and H. S. Maciel. 2015. Nanostructured thin films based on
TiO2 and/or SiC for use in photoelectrochemical cells: A review of the material characteristics, synthesis
and recent applications. Mater. Sci. Semicond. Proc. 29: 56–68.
Philippidis, N., E. Nikolakaki, S. Sotiropoulos, and I. Poulios. 2010. Photoelectrocatalytic inactivation of
E. coli XL-1 blue colonies in water. J. Chem. Technol. Biotechnol. 85: 1054–1060.
Price, M. J., and S. Maldonado. 2009. Macroporous n-GaP in nonaqueous regenerative photoelectrochemical
cell. J. Phys. Chem. C. 113: 11988–11994.
Radecka, M., M. Rekas, A. Trenczek-Zajac, and K. Zakrzewska. 2008. Importance of the band gap energy and
flat band potential for application of modified TiO2 photoanodes in water photolysis. J. Power Sources.
181: 46–55.
Rahman, M., B. H. Q. Dang, K. McDonnell, J. M. D. MacElroy, and D. P. Dowling. 2012. Effect of doping
(C or N) and co-doping (C + N) on the photoactive properties of magnetron sputtered titania coatings
for the application of solar water-splitting. J. Nanosci. Nanotechnol. 12: 4729–4735.
Rajeshwar, K., M. Kaneko, and A.Yamada. 1983. Regenerative photoelectrochemical cells using polymer-
coated n-GaAs photoanodes in contact with aqueous electrolytes. J. Electrochem. Soc. 130: 38–43.
52 Solar Energy Conversion and Storage

Ramaraj, R., and P. Natarajan. 1989. Photoelectrochemical investigations of phenosafranine dye bound to
some macromolecules. J. Chem. Soc. Faraday Trans. 1. 85: 813–827.
Redmond, G., D. Fitzmaurice, and M. Graetzel. 1994. Visible light sensitization by cis-bis(thiocyanato)
bis(2,2′-bipyridyl-4,4′-dicarboxylato) ruthenium(II) of a transparent nanocrystalline ZnO film prepared
by sol-gel techniques. Chem. Mater. 6: 686–691.
Rothenberger, G., D. Fitzmaurice, and M. Grätzel. 1992. Spectroscopy of conduction band electrons in trans-
parent metal oxide semiconductor films: Optical determination of the flatband potential of colloidal
titanium dioxide films. J. Phys. Chem. 96: 5983–5986.
Roušar, I., M. Rudolf, P. Lukášek, L. Kavan, N. Papageorgiou, and M. Grätzel. 1996. Optimization of param-
eters of an electrochemical photovoltaic regenerative solar cell. Sol. Energy Mater. Solar Cells. 43:
249–262.
Ryan, D. M., M. K. Coggins, J. J. Concepcion, D. L. Ashford, Z. Fang, L. Alibabaei, et al. 2014. Synthesis and
electrocatalytic water oxidation by electrode-bound helical peptide chromophore-catalyst assemblies.
Inorg. Chem. 53: 8120–8128.
Shi, J.-Y., W.-H. Leng, X.-F. Cheng, J.-Q. Zhang, and C.-N. Cao. 2005. Photocatalytic oxidation of methyl red
by TiO2 in a photoelectrochemical cell. Acta Phys. Chim. Sinica. 21: 971–976.
Singh, P., R. Singh, K. Rajeshwar, and J. DuBow. 1981. Photoelectrochemical behavior of n-GaAs elec-
trodes in ambient temperature molten salt electrolytes: Device characterization and loss mechanisms.
J. Electrochem. Soc. 128: 1145–1150.
Song, W., M. K. Brennaman, J. J. Concepcion, J. W. Jurss, P. G. Hoertz, H. Luo, et al. 2011. Interfacial electron
transfer dynamics for [Ru(bpy)2((4, 4′-PO3H2)2bpy)]2+ sensitized TiO2 in a dye-sensitized photoelectro-
synthesis cell: Factors influencing efficiency and dynamics. J. Phys. Chem. C. 115: 7081–7091.
Song, W., Z. Chen, C. R. K. Glasson, K. Hanson, H. Luo, M. R. Norris, et al. 2012. Interfacial dynamics and
solar fuel formation in dye-sensitized photoelectrosynthesis cells. Chem. Phys. Chem. 13: 2882–2890.
Song, W., A. Ito, R. A. Binstead, K. Hanson, H. Luo, M. K. Brennaman, et al. 2013a. Accumulation of multiple
oxidative equivalents at a single site by cross-surface electron transfer on TiO2. J. Am. Chem. Soc. 135:
11587–11594.
Song, W., H. Luo, K. Hanson, J. J. Concepcion, M. K. Brennaman, and T. J. Meyer. 2013b. Visualization of
cation diffusion at the TiO2 interface in dye sensitized photoelectrosynthesis cells (DSPEC). Energy
Environ. Sci. 6: 1240–1248.
Song, W., A. K. Vannucci, B. H. Farnum, A. M. Lapides, M. K. Brennaman, B. Kalanyan, et al. 2014. Visible
light driven benzyl alcohol dehydrogenation in a dye-sensitized photoelectrosynthesis cell. J. Am.
Chem. Soc. 136: 9773–9779.
Stergiopoulos, T., I. M. Arabatzis, M. Kalbac, I. Lukes, and P. Falaras. 2005. Incorporation of innovative com-
pounds in nanostructured photoelectrochemical cells. J. Mater. Proc. Technol. 161: 107–112.
Su, F., T. Wang, R. Lv, J. Zhang, P. Zhang, J. Lu, et al. 2013. Dendritic Au/TiO2 nanorod arrays for visible-light
driven photoelectrochemical water splitting. Nanoscale. 5: 9001–9009.
Sun, Y., A. C. Onicha, M. Myahkostupov, and F. N. Castellano. 2010. Viable alternative to N719 for dye-
sensitized solar cells. ACS Appl. Mater. Interfaces. 2: 2039–2045.
Sun, Y., K. Yan, G. Wang, W. Guo, and T. Ma. 2011. Effect of annealing temperature on the hydrogen produc-
tion of TiO2 nanotube arrays in a two-compartment photoelectrochemical cell. J. Phys. Chem. C. 115:
12844–12849.
Taratula, O., J. Rochford, P. Piotrowiak, E. Galoppini, R. A. Carlisle, and G. J. Meyer. 2006. Pyrene-
terminated phenylenethynylene rigid linkers anchored to metal oxide nanoparticles. J. Phys. Chem. B.
110: 15734–15741.
Thompson, D. W., A. Ito, and T. J. Meyer. 2013. [Ru(bpy)3]2+* and other remarkable metal-to-ligand charge
transfer (MLCT) excited states. Pure Appl. Chem. 85: 1257–1305.
Thompson, L., J. DuBow, and K. Rajeshwar. 1982. Photoelectrochemical generation of chlorine on catalyti-
cally modified n-silicon/indium tin oxide anodes. J. Electrochem. Soc. 129: 1934–1935.
Tributsch, H. 1980. Photoelectrochemical behaviour of layer-type transition metal dichalcogenides. Faraday
Discuss. Chem. Soc. 70: 189–205.
Vannucci, A. K., L. Alibabaei, M. D. Losego, J. J. Concepcion, B. Kalanyan, G. N. Parsons, et al. 2013.
Crossing the divide between homogeneous and heterogeneous catalysis in water oxidation. Proc. Natl.
Acad. Sci. USA. 110: 20918–20922.
Vlachopoulos, N., P. Liska, J. Augustynski, and M. Grätzel. 1988. Very efficient visible light energy harvest-
ing and conversion by spectral sensitization of high surface area polycrystalline titanium dioxide films.
J. Am. Chem. Soc. 110: 1216–1220.
Photoelectrochemical Cells 53

Vorobets, V. S., S. K. Kovach, and G. Y. Kolbasov. 2002. Distribution of ion species and formation of ion
pairs in concentrated polysulfide solutions in photoelectrochemical transducers. Russ. J. Appl. Chem.
75: 229–234.
Wadhawan, J. D., T. J. Davies, A. D. Clegg, N. S. Lawrence, J. C. Ball, O. V. Klymenko, et al. 2002.
Photoelectrochemistry of bromonitrobenzenes: Mechanism and photoelectrochemically-induced halex
reactions. J. Electroanal. Chem. 533: 33–70.
Wang, C., Z. Chen, H. Jin, C. Cao, J. Li, and Z. Mi. 2014. Enhancing visible-light photoelectrochemical water
splitting through transition-metal doped TiO2 nanorod arrays. J. Mater. Chem. A. 2: 17820–17827.
Wang, P., S. M. Zakeeruddin, P. Comte, I. Exnar, and M. Grätzel. 2003. Gelation of ionic liquid-based elec-
trolytes with silica nanoparticles for quasi-solid-state dye-sensitized solar cells. J. Am. Chem. Soc. 125:
1166–1167.
Watson, D. F., A. Marton, A. M. Stux, and G. J. Meyer. 2003. Insights into dye-sensitization of planar TiO2:
Evidence for involvement of a protonated surface state. J. Phys. Chem B. 107: 10971–10973.
White, J. R., F.-R. Fan, and A. J. Bard 1985. Semiconductor electrodes: LVI. Principles of multijunction elec-
trodes and photoelectrosynthesis at Texas Instruments’ p/n-Si solar arrays. J. Electrochem. Soc. 132:
544–550.
Wongwanwattana, P., P. Krongkitsiri, P. Limsuwan, and U. Tipparach. 2012. Fabrication and photocatalysis of
nanostructured TiO2 for solar hydrogen production. Ceramics Int. 38 (Suppl): S517–S519.
Xia, H.-L., F. Liu, S. Ardo, A. A. N. Sarjeant, and G. J. Meyer. 2010. Photoinduced electron transfer from Ru
am(m)ine compounds with low-lying ligand field excited states to nanocrystalline TiO2. J. Photochem.
Photobiol. A: Chem. 216: 94–103.
Xiang, C., G. M. Kimball, R. L. Grimm, B. S. Brunschwig, H. A. Atwater, and N. S. Lewis. 2011. 820 mV
open-circuit voltages from Cu2O/CH3CN junctions. Energy Environ. Sci. 4: 1311–1318.
Yang, J., D. Wang, H. Han, and C. Li. 2013. Roles of cocatalysts in photocatalysis and photoelectrocatalysis.
Acc. Chem. Res. 46: 1900–1909.
Yang, T., H. Wang, X.-M. Ou, C.-S. Lee, and X.-H. Zhang. 2012. Iodine-doped-poly(3,4-ethylenedioxythiophene)-
modified Si nanowire 1D core-shell arrays as an efficient photocatalyst for solar hydrogen generation.
Adv. Mater. 24: 6199–6203.
Yohannes, T., and O. Inganäs. 1999. All-solid-state photoelectrochemical energy conversion with the conju-
gated polymer poly[3-(4-octylphenyl)-2,2′-bithiophene](FT). Synth. Met. 107: 97–105.
Yujing X., Y. Sun, G. Wang, K. Yan, and J. Zhao. 2015. Effect of NH4F concentration and controlled-charge
consumption on the photocatalytic hydrogen generation of TiO2 nanotube arrays. Electrochim. Acta.
155: 312–320.
Zhao, Y., Y. He, D.-B. Xiong, W. Ran, Z. Liu, and F. Gao. 2014. Dual template synthesis and photoelectro-
chemical performance of 3-D hierarchical porous zinc oxide. Int. J. Hydrogen Energy. 39: 13486–13490.
5 Organic Photovoltaic Cells
Meenakshi Singh Solanki, Taruna Dangi, Paras Tak,
Sanyogita Sharma, and Rakshit Ameta

CONTENTS
5.1 Introduction............................................................................................................................. 55
5.2 History..................................................................................................................................... 57
5.3 Basic Processes........................................................................................................................ 58
5.3.1 Photon Absorption....................................................................................................... 58
5.3.2 Exciton Diffusion......................................................................................................... 58
5.3.3 Charge Separation....................................................................................................... 59
5.3.4 Charge Transfer........................................................................................................... 59
5.3.5 Charge Collection........................................................................................................60
5.4 Characteristics.........................................................................................................................60
5.5 Materials.................................................................................................................................. 61
5.5.1 Pigments...................................................................................................................... 62
5.5.2 Dyes............................................................................................................................. 62
5.5.3 Small Molecules, Oligomers, Polymers, and Dendrimers.......................................... 63
5.5.4 Liquid Crystals............................................................................................................ 65
5.5.5 Other Materials............................................................................................................66
5.6 Types of Organic Photovoltaic Cells....................................................................................... 68
5.6.1 Single Layer................................................................................................................. 68
5.6.2 Bilayer.......................................................................................................................... 69
5.6.3 Heterojunctions............................................................................................................ 71
5.7 Tandem Solar Cells.................................................................................................................. 75
5.8 Hybrid Tandem Photovoltaic Cell........................................................................................... 76
5.9 Inverted Tandem Solar Cell..................................................................................................... 77
5.10 Inverted Organic Photovoltaic Cells........................................................................................ 77
References......................................................................................................................................... 78

5.1 INTRODUCTION
There is a continuous increase in demand for electrical energy in almost all fields (i.e., from consumer
electronics, to small-scale distributed power systems, to centralized megawatt-scale power plants, in
domestic use, for transportation, etc.). Conventional methods for generating electrical energy are time
consuming, high in cost, low power generating, and release harmful by-products into the environ-
ment, which lead to global warming. Global warming is clear evidence of the fundamental idea of
Newton that there is no action without reaction. But renewable energies produced from our natural
environment, such as wind, solar, thermal, photovoltaic, geothermal, marine, and hydropower, help
us to reduce our dependence on conventional methods. Therefore, solar energy, a clean alternative to
traditional methods of power generation for sustainable development, is a good as well as promising
choice. Direct utilization of solar radiation is similar to the ideal use of solar radiation in nature in the
form of photosynthesis.
Solar electricity is a steadily growing energy technology. Among different technologies
already available to convert solar light directly into electricity, photovoltaics (PVs) offer several

55
56 Solar Energy Conversion and Storage

benefits. Photovoltaic cells can produce power near the end user of the electricity, thus avoiding
transmission losses and costs. Solar panels do not produce any noise or greenhouse gas emis-
sions and require very little maintenance. Photovoltaic devices based on inorganic materials
have commonly been used, but photovoltaic cells based on organic materials are garnering
attention. The main advantages of organic photovoltaic (OPV) cells compared to other types of
photovoltaic cells are their low weight, attractive form factor, scalability, flexibility, and low-
cost fabrication.
The term photovoltaic is derived from photo meaning “light” and voltaic meaning “electricity”;
thus, photovoltaic cells are cells that convert sunlight directly into electricity at the atomic level.
Photovoltaic cells are made of special materials called semiconductors, such as silicon, that exhibit
a property known as the photoelectric effect. Basically, when light energy strikes the cell, loosely
bound electrons are knocked out from the atoms in the semiconductor material. When these free
electrons are captured, the result is an electric current that can be used as electricity.
Silicon-based photovoltaic cells account for the major portion of production of solar electricity
in the world. However, although the prices of these cells have been drastically reduced, they are
still too expensive. Therefore, organic compounds are used to overcome such high cost problem.
Organic semiconductors are a comparatively less expensive alternative to inorganic semiconduc-
tors. Moreover, they are less energy consuming. One of the challenges in realizing efficient organic
photoconversion systems (organic PV cells) is that the electron–hole pair created via photon absorp-
tion must overcome the Coulomb attraction (i.e., losing the strongly bound electron and hole that
result from photon exposure) to generate photocurrent. Many compelling solutions to this problem
have been proposed.
The main difference between organic and inorganic materials is that in an inorganic semicon-
ductor, the free charge carriers, electrons, and holes are created directly upon light absorption,
whereas electrostatically bound charge carriers, excitons, are formed in an organic semiconductor:

• The dielectric constant and Bohr radius of carriers are less in organic semiconductors as
compared to inorganic semiconductors.
• An organic semiconductor has an easy processing at 20°C–200°C, but on the contrary, an
inorganic semiconductor requires high temperatures of 400°C–1400°C.

The main benefit of an organic cell over an electrochemical cell is the absence of a liquid electrolyte.
The active layer thickness of an organic solar cell is only 100 nm, which is 1000 times thinner than
that of a Si-solar cell and 10 times thinner than that of an inorganic thin-film solar cell. Therefore,
organic solar cells have the potential for cost-efficient and large-scale applications.
Organic photovoltaics, a solar cell technology, is related to the idea of providing flexible, wear-
able, lightweight, and low-cost photovoltaic materials such as polymers (plastics), dyes, or certain
organic electron donor and acceptor molecules. In the last few decades, this field of organic pho-
tovoltaic materials has attracted scientific and economic interest, which is triggered by a rapid
increase in power conversion efficiencies. Commercially available organic solar cells have power
conversion efficiencies less than 3%, while silicon-based solar cells have power conversion effi-
ciencies of 40.7% on average. Organic photovoltaic devices have not yet reached the limits of their
inorganic counterparts (i.e., approximately 10%–20%) (Harald and Sariciftci 2004). Organic solar
cells are developing in a dynamic way, as they have low production costs because they are printed
onto a substrate. However, silicon-based solar cell technology has the limitation of extremely high
manufacturing and material costs.
Tang (1986) has reported a bilayer thin-film organic photovoltaic cell of copper phthalocyanine
and a perylene tetracarboxylic derivative with efficiency up to 1%. This lower yield (almost 1%) has
now increased almost 10 times in some other OPV cells. A photovoltaic conversion efficiency of
about 5% was observed in a solar cell having ZnO and pentacene-based single crystals and thin-
film heterojunction devices (Schon et al. 2000). However, the major objective in this field remains
Organic Photovoltaic Cells 57

the cost reduction of photovoltaic modules. Thin-film technology or silicon solar cells are still more
efficient, but there is a pressing demand to develop technology that is eco-friendly, is readily avail-
able, and utilizes low-cost materials. Polymeric materials, particularly plastics, have great potential
in these aspects.

5.2 HISTORY
Bequerel (1839) was the first to notice the photoelectric effect. He found that platinum electrode
coated with silver halogen produces small amounts of electric current when irradiated with elec-
tromagnetic radiation in aqueous solution. Smith (1873) discovered the photoconductivity of sele-
nium. The next significant photovoltaic development was the photoconductive effect in selenium
placed between two metal electrodes (Adams and Day 1877). Adams and Day observed an anom-
aly that could be explained by the generation of internal voltages. Heated platinum contacts were
pushed into opposite ends of small cylinders of vitreous selenium. They also reported that it was
possible to get current in the selenium by the action of light. This was the first demonstration of
the photovoltaic effect in a solid-state system. The photogenerated currents were attributed to
light-induced crystallization of the outer layers of the selenium bar. The photovoltaic action of the
selenium differed from its photoconductive action, where a current was produced spontaneously
by the action of light.
The next significant step was taken by Fritts (1883). He coated molten selenium plate with an
extremely thin layer of gold to prepare the first thin-film photovoltaic device, which was as large as
30 cm2 in area, but power conversion efficiency was limited to less than 1%. The photoconductivity
in an organic compound (anthracene) was reported by Pochettino (1906).
Grondahl (1933) reported photovoltaic cells based on a copper-cuprous oxide junction. In this
model, Pb wire was used to provide firm contact to the illuminated surface of the cell. Bergmann
(1931) reported a selenium device that was much more effective than a copper-based device and
became the commercially dominant product. A thallous-sulfide cell of similar performance was
also studied (Nix and Treptwo 1939). Excited by this outstanding achievement, the New York Times
forecasted that solar cells would eventually harness the “limitless energy of the sun.” In 1958, PV
array-powered radios appeared on the U.S. Vanguard I space satellite, and this was the first time
that PV technology was practically utilized. During this period, PV cells made of cadmium sul-
fide, gallium arsenide, cadmium telluride, and indium phosphide also appeared. However, each
technology had its own disadvantages, as well as merits. Cadmium is used in CdS and CdTe PV
cells.
These early PV cells of copper-cuprous oxide thin film, lead sulfide, and thallium sulfide were
thin-film Schottky barrier devices, where a semitransparent layer of metal deposited on top of the
semiconductor provided both asymmetric electronic junctions, which are necessary for photovoltaic
action, and access to the junction for the incident light.
However, photoconductivity was what excited researchers, not the PV properties of materials
like selenium. The current generated was proportional to the intensity of the incident light, and this
was related to the wavelength. The photoconductive materials were ideal for photographic light
meters. Barrier structures in the photovoltaic effect were an added benefit, which means the light
meter could operate without a power supply. It was not until the 1950s that potentially useful quan-
tities of power were produced by PV devices in crystalline silicon by development of good-quality
silicon wafers. These were used in the new solid-state electronics devices.
In 1954, Bell Labs revealed the first high-power silicon PV cell using a p-n junction with power
conversion efficiency (PCE) of 6%. Photoconductivity in the poly(N-vinyl-carbazole) (PVK) poly-
mers was discovered by Hoegel (1965). The concepts for organic semiconductors were established
by Chiang et al. (1977) and Shirakawa et al. (1977). Their findings were astonishing, because they
highlighted the potential transition of photovoltaic substances from inorganic to organic semicon-
ductors, which could lower production costs and ease of processing.
58 Solar Energy Conversion and Storage

In the last few decades, scientists have had great interest in developing organic PV devices.
During the 1990s, it was shown that the quantum efficiency of electron transfer is very high in
excited polymers and C60, which is promising for change carrier separation in PV cells (Sariciftci
et al. 1992, 1993). Organic displays with organic light-emitting devices (OLEDs) have introduced
new technology (Diaz et al. 2001).

5.3  BASIC PROCESSES


Typically, a photovoltaic cell is made up of two photoactive materials placed in between two metal-
lic electrodes, including a transparent electrode, which is exposed to light. Photons with sufficient
energy are adsorbed by the surface, and loosely bound electrons are removed. After the charge
separation process, the charge carriers (electrons) are transported to the electrodes. Finally, charges
are collected by the electrode, and these will produce electric current in the external circuit. These
steps are important steps of any photovoltaic process (Chamberlain 1983). Basically, there are five
important steps to achieve high conversion efficiency of solar radiation into electrical energy:

• Photon absorption
• Exciton diffusion
• Charge separation
• Charge transfer
• Charge collection

5.3.1  Photon Absorption


The first process in any photovoltaic device is the absorption of photons. A photosensitive semicon-
ductor is exposed to light, leading to excitation of electrons from the highest occupied molecular
orbital (HOMO) to the lowest unoccupied molecular orbital (LUMO). It may be easier for any
inorganic photovoltaic devices to have better absorption, but it is slightly difficult for organic photo-
voltaic devices, because the bandgap is quite high for organic materials. The bandgap of an organic
material should be 1.1 eV to permit sufficient absorption of solar radiation reaching Earth (about
80%). This corresponds to 1100 nm, while most of the semiconducting polymers have bandgaps
more than 2.0 eV, corresponding to 600 nm. This limits the absorption of solar radiation to about
30%. Normally, the organic layer is thin, but in photovoltaic devices, the low charge carrier and
exciton mobilities require slightly thick layers (~100 nm). As the absorption coefficient of organic
materials is relatively higher than that of silicon, layers of 100 nm thickness can absorb 60%–90%
solar radiation provided that a reflective back contact is used. Therefore, in order to increase light
absorption, efforts are being made to develop small-bandgap polymers.
Ko et al. (2011) reported a threefold enhancement in photon absorption in specific regions of the
solar spectrum, in part through multiple excitation resonances by the two-dimensional (2D) pho-
tonic crystal (PC) geometry in comparison to the geometry of a conventional planar cell. Photonic
crystal geometry is developed using a material-agnostic process called PRINT (pattern replication
in nonwetting templates) on an organic photoactive bulk heterojunction blend of P3HT (poly(3-
hexylthiophene)) and PCBM (phenyl-C61-butyric acid methyl ester). In this process, highly ordered
arrays of nanoscale features were readily made in a single processing step over wide areas (4 cm2)
that are scalable, exhibiting efficiency improvements of 70% that result not only from greater
absorption, but also from electrical enhancements.

5.3.2  Exciton Diffusion


Excitons are generated on photon absorption. These excited excitons should reach a dissociation site.
The diffusion length should be at least equal to the thickness of the layer for sufficient absorption,
Organic Photovoltaic Cells 59

as such sites are normally available at the other ends of the semiconductors. If this is not the case,
then the excited excitons may recombine, and photons are wasted. The diffusion ranges in pigment
and polymers are ordinarily quite low (~10 nm).

5.3.3 Charge Separation
Electrons are transferred to an electron acceptor from electron donor material in the charge separa-
tion process. The charge separation may occur at organic semiconductor/metal interfaces at oxygen
as impurity or it may be between materials with quite different electron affinities and ionization
potentials. If the difference between ionization potential and electron affinity is not sufficient, then
excitons may just jump on the material with the lower bandgap without splitting its charges. As a
result, it may recombine and not contribute changes to the photocurrent. An electron donor material
has small electron affinity. On the contrary, an electron-accepting material has high electron affin-
ity. This difference in electron affinity is the driving force that leads to exciton dissociation.
It is very important for an efficient charge generation that most of the energy of the absorbed
photon is used for the charge separation process. This means that the energy of absorbed light
should not be wasted in competitive processes (fluorescence or nonradiative decay). The separated
charge should be stabilized in order to transfer photogenerated charge to one of the electrodes (i.e.,
the back-electron transfer or recombination should be minimized).
Excitons must overcome the Coulomb attraction force present between them for long-range
charge separation. The formation of an excited, delocalized band state helps to obtain long-range
charge separation, which has a positive effect on the efficiency of photovoltaic devices. Exposing
excitons under infrared radiation (photon) for less than 1 ps results in the transfer of a charge-pair
bound at a heterojunction to delocalized band states, which act as the gateway for charge separation
(Bakulin et al. 2012).
The charge separation mechanism helps in increasing the performance of organic solar devices.
Gelinas et al. (2014) studied the ultrafast long-range charge separation process. They found that
charge separations of an electron–hole pair generated by light absorption across the donor–acceptor
heterojunction are time dependent in organic solar cells. They also found that when charge sepa-
ration distance is near 4 nm, then about 200 meV of electrostatic energy is created from charge
separation within 40 fs of excitation. At this stage, the residual Coulomb attraction between charges
is at or below thermal energies, so charges (electron and hole) are separated completely. This early
behavior is consistent with charge separation through access to delocalized π-electron states in
ordered regions of the fullerene acceptor material.

5.3.4 Charge Transfer
The transport of charges is affected by recombination as it moves to the electrode. This is more
specific when the same material is used as the transport medium for both carriers (i.e., electrons
and holes). Their interaction with other atoms or charges will lower the speed and, consequently,
limit the current. Charge transport is also low due to less intrinsic mobility in organic material and
by the charge trapping effects of impurities and defects. The interfacial charge pair resides at the
donor–acceptor heterointerface, which is called the charge-transfer complex. It is responsible for
the photocurrent and the open-circuit voltage (Deibel et al. 2010).
The dissociation of excitons into free electrons and holes at donor–acceptor heterointerfaces
leads to photocurrent generation in OPV cells. It was also observed that the dielectric constant helps
in determining the charge-transfer energy. A low dielectric constant results in strong Coulomb
interactions between electron–hole pairs, which opposes the generation of free charges. The charge-
transfer polarizability increases when the C60 crystallite size exceeds the threshold (i.e., approxi-
mately 4 nm). The charge-transfer step plays a major role in the performance of OPV cells (Piliegoa
and Loi 2012).
60 Solar Energy Conversion and Storage

Jailaubekov et al. (2013) used femtosecond nonlinear optical spectroscopies and nonadiabatic
mixed quantum mechanics/molecular mechanics simulations in the phthalocyanine-fullerene
model OPV system to explain hot charge-transfer exciton formation and relaxation. The first excita-
tion on phthalocyanine produces hot charge-transfer excitons in 10–13 s and then relaxation to lower
energies and shorter electron–hole distances of 10–12 s. Thus, this hot CT exciton cooling process
and collapse of charge separation set the fundamental time limit for competitive charge separation
channels, which leads to efficient photocurrent generation. Bernardo et al. (2014) investigated the
charge-transfer process, which is affected by variation of fullerene fraction in small molecule–
fullerene bulk heterojunction PV cells.
As organic solids have lower electron mobility, inorganic materials are also used as electron-
transporting components. TiO2 has been used to increase performance of PVs such as conjugative
polymer, poly(p-phenylene vinylene) (PPV) with ultrathin TiO2 nanocrystal (Salafsky 1999), and
mesoporous titania (Coakley and McGehee 2003). Charge separation and charge transport are dif-
ficult in some organic materials like polymers. Elongated crystalline components are attractive as
electron transporters if crystal size and orientation can be managed. The charge-transfer and charge
separation processes in PV cells are increased by a soluble perpylene dye–polymer blend (Dittmer
et al. 2000) and graphene (Chang et al. 2014).

5.3.5 Charge Collection
In most organic solar cells, the separated electrons and holes are transported to the opposite elec-
trodes in an internal electric field created by asymmetry of the electrodes (different work functions)
or in built-in potentials. Selective doping enhances the performance of the PV cell as it provides
low serial resistance and creates internal electric fields for the collection of charges. The multilayer
model of ITO/poly(2,5-dioctyloxy-phenylene vinylene) (OOPPV)/octaethylporphine (OEP)/C60/Al
has double heterojunction of OOPPV/OEP and OEP/C60, which leads to charge generation by exci-
tonic dissociation (Yoshino et al. 1997).
Good nonblocking contacts between the molecular materials and the electrodes of the cell are
needed for efficient charge collection. An additional material layer between the metal and organic
layer is also used to facilitate good ohmic contact in some cases, such as with LiF between conju-
gated polymer/methanofullerene (MDMO-PPV-PCMB) blend and aluminum (Brabec et al. 2001a;
Shaheen et al. 2001a).
Charge collection is affected by the presence of inorganic film as the electron and hole collection
layer. Poly(3-hexylthiophene (P3HT) and [6,6]-phenyl-C71-butyric acid methyl ester (PC71BM) as the
electron donor and electron acceptor, respectively, were prepared by Vasilopoulou et al. (2014a). They
used ZnO film as the electron-collecting layer and an under-stoichiometric molybdenum oxide MoOx
as the hole collection layer to improve the inverted organic PV cell efficiency. The ZnO film was
prepared using two methods in this fabrication: by atomic layer deposition (ALD-ZnO) and by using
the sol-gel method (sg-ZnO). Both films had the same thickness of 20 nm, but results proved that the
ALD method shows significantly enhanced efficiency as it develops conformal and defect-free ZnO
electron collection layers, which are responsible for high-performance organic photovoltaics.

5.4 CHARACTERISTICS
Some characteristics of the solar cell that determine its photovoltaic performance and electrical
behavior are

• Open-circuit voltage (Voc): This is maximum voltage across the PV cell in solar radiation,
when no current is flowing in the PV cell.
• Short-circuit current (isc): This is the current that flows through an exposed PV cell when
there is no external resistance. The maximum current that a device is able to produce is
Organic Photovoltaic Cells 61

called the short-circuit current. Under an external load, the current will always be less
than isc.
• Fill factor (FF): This is the ratio of actual maximum power output to its theoretical power
output (maximum current isc and voltage Voc). It is the key quantity used to measure per-
formance of the cell:

Pmax V ×i
FF = = max max (5.1)
Voc × isc Voc × isc

where imax and Vmax are maximum current and potential, respectively, and Pmax is their product.
• Quantum efficiency (QE): This is the efficiency of a PV cell as a function of the energy or
wavelength of the incident radiation. It specifically relates the number of charge carriers
collected to the number of photons shining on the device for a particular wavelength. There
are two types of quantum efficiency:
• External quantum efficiency: This involves the losses by reflection and transmission.
It is also known as incident photon-to-current efficiency (IPCE):

J sc × h × c
Incident photon-to-current efficiency ( IPCE ) = (5.2)
I × λ × e 

where I is the intensity of incident radiation (W cm−2), λ is wavelength (μm), and Jsc is
short-circuit current density (A cm−2), while h, c, and e have their usual significance as
Planck constant, velocity of light, and electronic charge, respectively.
The formula is thus

1.24J sc
IPCE = (5.3)
I ×λ

As h, c, and e are constant, the external photovoltaic yield (η) is the ratio of Pmax to the
product of I and surface (S) of the modules:

Pmax
η= (5.4)
I ×S
• Internal quantum efficiency: This includes losses because of reflection and transmis-
sion of light energy such that it considers processes involving absorbed photons only.
By accounting for transmission and reflection processes, external quantum efficiency
can be transformed into internal quantum efficiency.
• Power conversion efficiency (PCE, η): This is the ratio of power output to power input
(i.e., the amount of power generated by the PV cell compared to the power available in
the incident photons (Pin).

5.5 MATERIALS
Inorganic materials have been commonly used in photovoltaic applications, but there are certain
added advantages in using organic materials for this purpose. The main advantages of using organic
materials for photovoltaic application are

• Organic materials have a high optical absorption coefficient, so a large amount of light can
be absorbed with a small amount of material. Therefore, the amount of organic materials
required is comparatively smaller than the amount of inorganic materials required.
62 Solar Energy Conversion and Storage

• Organic materials can be produced more easily than inorganic materials.


• Organic materials can be processed easily using wet processing (spin coating or doctor
blade technique) or dry processing (evaporation via mask).
• Chemical/physical properties, such as solubility, valence, bandgap, and charge transfer,
can be easily adjusted in the case of organic materials.
• A large variety of functionalities of the organic materials can be used, such as dyes, pig-
ments, oligomers, polymers, dendrimers, and liquid crystals.

Organic photovoltaic materials mainly consist of three different categories, based on their mechani-
cal or processing properties: insoluble, soluble, or liquid crystalline. They may be further catego-
rized on the basis of conjugated π electrons such as molecules with single structural repeat units
(monomer), few structural repeat units (oligomers), and more than about 10 repeat units (poly-
mers). Oligomers and monomers that absorb in the visible light are known as chromophores and are
referred to as dyes, if they are soluble, or pigments, if they are not (Petritsch et al. 2000). Basically,
organic solar cells, which belong to the class of photovoltaic cells, are known as excitonic solar cells
(Thompson and Fréchet 2008).
The main disadvantages related to organic materials are their low efficiency, strength, and
stability.

5.5.1  Pigments
Pigments are used for coloring a material. They form polycrystalline thin film on evaporation (sub-
limation) from solid powder. Perylene, or perylenetetracarboxylic acid diimide, and phthalocya-
nine, or different metallophthalocyanines, give the structural backbone to many molecules used in
efficient pigment- and dye-based organic solar cells. Fullerene (C60) and pentacene are insoluble in
most of the solvents; therefore, these are considered as pigments.

5.5.2  Dyes
Dyes impart color to any material on attachment. Their thin film cannot be obtained so easily from
solvent, but they can be easily incorporated into a polymer host (Gautier-Thianche et al. 1998; Kido
et al. 1994). Dyes can also be attached chemically to the polymer host (Cacialli et al. 1998; Jiang
et al. 2000).
Photoconductivity in anthracene was observed in the beginning of the twentieth century
(Pochettino 1906). Until the 1980s, the performance of a dye-sensitized solar cell was about 0.1%.
A major breakthrough came when Tang (1986) showed that much higher efficiencies are attainable
by producing a double-layered cell using two different dyes.
Organic dyes were also used for sensitization of solar cells. Basically, a monolayer of a dye is
introduced between the donor and the acceptor to increase the absorption of light. Yoshino et al.
(1997) studied a photovoltaic device made up of the three organic layers (i.e., donor–absorber–
acceptor structure) to enhance photovoltaic efficiency. For a wider absorption band, there is an
alternate route that replaces the electron-transporting polymer with a polymer blend with crystalline
dyes. Dye crystals like anthracene or perylene were used as electron acceptors in polythiophene-
based solar cells (Dittmer et al. 2000). The nature of the dye-sensitized solar cell was explained by
Cahen et al. (2000).
Fanshun and Tian (2005) reported the potential use of cyanine dyes in the thin-film hetero-
junction of organic photovoltaic devices because it widens the absorption spectra and photoaction
spectra in the visible region and, therefore, increases the performance of the cell. Walter et al.
(2010) utilized organic dyes porphyrins, phthalocyanines, and related compounds as components
of solar cells, including organic molecular solar cells, polymer cells, and dye-sensitized solar cells.
Chen et al. (2013) reported the use of ternary organic photovoltaic systems, including polymer/
Organic Photovoltaic Cells 63

small molecule/functional fullerene, polymer/polymer/functional fullerene, small molecule/small


molecule/functional fullerene, polymer/functional fullerene I/functional fullerene II, and polymer/
quantum dot or metal/functional fullerene systems in organic photovoltaic cells. Kim et al. (2014a)
were able to increase PCE to 3.08% and open-circuit voltage to 1.03 V by introducing carbazole
or fluorene-containing rhodanine dyes as acceptors in OPV cells. The absorbing layer is polarized
upon light absorption to drive the charges toward the appropriate transport layers.

5.5.3 Small Molecules, Oligomers, Polymers, and Dendrimers


Oligomers are typically formed during polymerization, when the number of repeat units is 2–12.
The oligomer has the desired electrical and optical properties of a polymer, and their thin films are
usually available in a polycrystalline state. If the repeat units are more than the limit of oligomer,
then it is called a polymer. When the similar repeat units are present in three-dimensional (3D)
form, it forms dendrimers. These dendrimers may have different shapes, such as branches of a star
or an ellipsoid-like structure (Halim et al. 1999).
Conjugated polymers are attractive semiconductors for photovoltaic cells because they have large
conjugative system and low bandgaps to broaden the absorption range; therefore, they are strong
absorbers and can be deposited on flexible substrates at low cost. Conjugative polymers are organic
molecules consisting of repeating structural units attached to each other by alternating carbon–carbon
single and double bonds. Such polymers are made conductive by doping. These can be oxidized or
reduced to produce p-doped or n-doped materials, respectively. The p-doped polymers are utilized
in many devices, such as electrochromic devices, rechargeable batteries, capacitors, and membranes,
and n-doped materials have not garnered as much attention. On the basis of molecular structure and
chemical composition, polymers can be used as either electron donor or electron acceptor materials in
the organic solar cells. Some common semiconducting polymers, fullerene and its derivatives are as
presented in Figure 5.1.
Plastic is an insulator, but it was made a conductor by adding some impurity, and the conductivity
of the polymer increased more than a billion times. This discovery and development of electrical
conductive polymers snatched a Nobel Prize for Heeger, McDiarmid, and Shirakawa. Chiang et al.
(1977) reported an increase in conductivity of doped polyacetylene. Shirakawa et al. (1977) treated
polyacetylene with an electron acceptor such as iodine and observed higher conductivity (of the
order of 1010 times) than the pure acetylene.
One of the most studied photoconducting polymers is poly(vinyl carbazole) (PVK). Photoconductivity
of PVK was also increased by doping with fluorinated styrene chromophore (Hendrickx et al. 1999)
and CdS nanocrystals (Wang et al. 2000). The conjugative polymers exhibit absorption in the blue or
green region. New material combinations like polymer–fullerene were synthesized, which lower the
bandgap and show absorption in the red or infrared region. The polymer–fullerene was used (Yu et al.
1995) in place of poly(para-phenylene vinylene) (PPV) (Jenekhe and Yi 2000), polythiophene and its
derivatives (Brabec et al. 2001b; Too et al. 2001), polypyrrole/thiazadole copolymers (Dhanabalan
et al. 2001), and thiophene/2-isothianaphthene copolymers (Shaheen et al. 2001b). Recent develop-
ment of organic conjugated polymer–based photovoltaic elements has helped us in achieving low-cost
and easy-to-produce energy from light.
The plastic solar cell (PSC) or polymer solar cell uses conductive organic polymers or small
organic molecules for absorption of light and charge transport to generate electricity from sunlight
(Nelson 2002). The photoactive device is based on the photoinduced charge transfer from donor-
type semiconducting polymers onto acceptor-type polymers such as Buckminsterfullerene; this is
the fundamental process used in photovoltaic devices. Most of the conjugated polymer can be uti-
lized as an electron donor because on photoexcitation, electrons are promoted to the antibonding
band. As the photoexcited electron is transferred to an acceptor unit, the resulting cation radical
(positive polaron) species on the conjugated polymer backbone is known to be highly delocalized,
mobile, and stable.
64 Solar Energy Conversion and Storage

CH 3

CH3

C H
C H C H
3
3 3

n
CH
3

CH CH

3 3

n n
O
Polyacetylene Poly(phenylene vinylene) (PPV) CH3
MEH-PPV

OCH3

O
OC6H13 CH 3

CH
3

OC6H13
H3C CN

C6H13O NC CH
3

n
C6H13O

CN-PPV Fullerene PC61BM

FIGURE 5.1  Some molecules used in OPV cells.

Charge carrier mobilities can be as low as 10 –4  cm2/V⋅s (approximately) in polymeric semi-
conductors. Because of the limit on increasing the thickness of the photoactive layer, the series
resistance becomes dominant, and the short-circuit current breaks down. Using poly(p-phenylene
vinylene) (PPV) with fullerenes, one can obtain an optimum thickness of ~100  nm. Therefore,
instead of increasing the thickness of the photoactive layer, higher charge carrier mobility materi-
als are required. Regioregular poly(3-alkylthiophenes) (P3AT) have better mobilities and produce
higher photocurrents by increasing the thickness of the layer. Charge mobility is also sensitive to
solid-state nanoscale morphology, which can be improved by different methods such as changing
the solvent of the casting solution, as well as tempering the cast films. Bandgap engineering will
also help in increasing the performance of the semiconducting polymer. It means different-colored
semiconductor polymer can be achieved by modifying the chemical structure, which leads to a
change in bandgap.
Coakley and McGehee (2003) prepared the conjugated polymer regioregular poly(3-hexylthiophene)
into films of mesoporous titania by infiltration. The mesoporous titania films have pores with diam-
eter less than 10 nm, which provide regular pathways for electrons to travel to an electrode after elec-
tron transfer. A 1.5% power conversion efficiency and an external quantum efficiency of 10% under
monochromatic 514 nm light were shown. They also described that on blending polymers with electron-
accepting materials such as C60 derivatives, cadmium selenide, and titanium dioxide, power conversion
efficiencies reached 4% (Coakley and McGehee 2004).
The open-circuit voltage, which depends on the LUMO of the acceptor as well as on the HOMO
of the donor, is influenced by a donor–acceptor pair and wave function of electrode. The dependence
of the charge transport levels on temperature and light intensity also affect open-circuit voltage
values. At normal room temperature, it is approximately 0.8–0.9  V for poly(2-methoxy-5-(3′,7′-
dimethyloctyloxy)-1,4-phenylene vinylene) (MDMO-PPV) copolymer blended with [6,6]-phenyl-
C61 butyric acid methyl ester (PC61BM) and 0.5–0.6 V for poly-(3-hexylthiophene) (P3HT) blends
Organic Photovoltaic Cells 65

with PCBM (Sariciftci 2004). Exposing an organic layer of an organic photovoltaic device to solvent
vapor also affects its performance. Miller et al. (2008) reported an improvement in the performance
of organic poly(3-hexylthiophene) and phenyl-C61-butyric acid methyl ester devices via room tem-
perature solvent vapor annealing.
A new conjugative polymer semiconductor was obtained by combining tetracyclic lactum mono-
mer with thiophene (PTNT), which is a broad bandgap semicrystalline polymer with bandgap
2.2 eV. Such cells exhibited a fill factor around 0.6, high open-circuit voltage of 0.9 V, and a power
conversion efficiency of 5% for more than 200 nm (or ~ 400 nm) thick active layers (Kroon et al.
2014). These values are on the higher side of those reported for a conjugated polymer with broad
bandgap. Therefore, a newly developed tetracyclic unit may find some interesting applications in the
future, mainly for ternary tandem PVs.
Organic photovoltaic cells must be designed in such a way that a weak electron-donating unit is
used with a strong electron-withdrawing unit in order to decrease the HOMO energy level. Therefore,
a conjugative polymer must be attached with a pull-and-push unit for better photon adsorption.
The benzo[1,2-b:4,5-b′]dithiophene is one of the most used and effective push units. It is used with
pull units such as thieno[3,4-b]thiophene (TT) and benzo[2,1,3]thiodazole (BT) for form pull–push
copolymer PBDTTT and PBDT-DTBT, respectively. Various organic solar devices with pull-and-push
units, with their power conversion efficiency (%), are poly(3-(2′-methoxy-5′-octylphenyl) thiophene)
(POMeOPT)-poly(2,5,2′,5′-tetrahexyloxy-7,8′-dicyanodi-p-phenylene vinylene) (CN-PPV), 4.5
(Roman et al. 2003); 2,5-di(thiophen-2-yl)thieno[3,2-b]thiophene and thieno[3,4-c] pyrrole-4,6-dione
(PDTTTPD)-[6,6]-phenyl-C61-butyric acid methyl ester (PC61BM), 5.1 (Chen et al. 2011a); benzo[1,2-
b:4,5-b′]dithiophene (BDT)-5,6-bis(octyloxy)benzo[c] [1,2,5]oxadiazole (BO) units, 5.7 (Jiang et al.
2011); N-alkylthieno[3,4-c]pyrrole-4,6-dione (TPD) and PC61BM, 6.8 (Piliego et al. 2010); [6,6]-phe-
nyl-C71-butyric acid methyl ester (PBDTTT- PC71BM), 9.2 (He et al. 2012); and so on.
The two naphthalene diimide (NDI) dimers, bis-NDI-T-EG and bis-NDI-BDTEG, bridged
by thiophene and benzodithiophene, respectively, and symmetrically substituted by 2-methoxy-
lethoxyl in the bay region were synthesized. These two NDI dimers exhibit broad absorption in
the visible region of 300–650 (800) nm and display a HOMO/LUMO energy level of 5.88/3.80
and 5.46/3.78 eV, respectively. The devices with PBDTTT-C-T as donor material exhibit the best
efficiency for both bis-NDI-T-EG and bis-NDI-BDT-EG of 1.31% and 1.24%, respectively (Wang
et al. 2014a).
Organic materials drew attention because of their potential in providing environmentally safe,
flexible, lightweight, and inexpensive electronics. The processing cost is also reduced, if the poly-
mer is soluble. Linear polymers such as polyacetylene are much less soluble.

5.5.4 Liquid Crystals
Liquid crystals have emerged as a new category of organic solar cell materials. Their main feature
is that they provide order, permitting different properties such as light absorption, charge genera-
tion, and most importantly, high charge carrier mobility and long exciton diffusion lengths (a few
100 nm), which could be beneficial for organic solar cells (Petritsch et al. 1999; Seguy et al. 2000).
These materials show the phase with properties somewhere in between those of liquids and solids
at certain temperatures. The liquid crystalline molecules are arranged in such a way that they look
like crystalline solids but exhibit the mechanical properties of liquids (i.e., they are soft). Liquid
crystals may be dyes, pigments, oligomers, or polymers. They can be processed by both wet and dry
methods. They can have different shapes (smectic, nematic, columnar, etc.).
Langmuir-Blodgett films (Cimrova et al. 1996; Tokuhisa et al. 1998; Wu et al. 1996) and self-
assembled monolayers (Appleyard et al. 2000) also permit molecular order control similar to that of
liquid crystals. Therefore, they are used to adjust the properties of an electride (Nuesch et al. 1997).
A self-organized liquid crystal organic solar cell was reported by Schmidt-Mende et al. (2001),
using bilayer of liquid crystalline hexaphenyl-substituted hexabenzocoronene (HBCPhC12) as an
66 Solar Energy Conversion and Storage

electron donor and a perylene dicarboxylic acid diimide derivative as an electron acceptor in the
active layer of the cells with high external quantum efficiency of over 34% at 490 nm.

5.5.5 Other Materials
Metals with low work function are typically used in organic solar devices because these metals
are used to control carrier selectivity, transport, extraction, and blocking, as well as interface band
bending. But there are some disadvantages to these materials, including that they are generally
prone to reactions with water, oxygen, nitrogen, and carbon dioxide from air, leading to rapid device
degradation. Therefore, lanthanides were used as new metallic cathode interlayer materials that
increase device stability and still provide device efficiency similar to that achieved with a Ca inter-
layer (Nikiforov et al. 2013).
Chauhan et al. (2014) reported the performance of an organic solar cell with Al-doped ZnO thin
films (AZO films) prepared by radio frequency (RF) sputtering at different argon pressures and low
substrate temperatures of 80°C–95°C. As argon pressure was increased, their morphology, optical
absorption, photoluminescence spectrum, and electrical behavior also changed. At 0.15 Pa argon
pressure, Al-doped ZnO thin films show a Wurtzite-type hexagonal structure with [0001] preferred
orientation, high optical transmittance of ∼85% in the wavelength range 400–800  nm, and low
electrical resistivity of 9.54 × 10−4 Ω cm.
Current generation in an organic photovoltaic cell mainly depends on the micro- and nanostruc-
tures in the semiconductor layers. Gilchrist et al. (2014) used high-resolution transmission electron
microscopy (HRTEM) to gain unprecedented insights into the structure and composition of the
molecular layers within the depth of the device structure. The technique was applied to a solar cell
made of copper phthalocyanine (CuPc) and C60.
Nielsen et al. (2014) developed electron-deficient truxenone derivatives as an alternative to fuller-
ene. These derivatives are far better than PCBM because they have easily tunable absorption pro-
files, higher electron affinities, higher absorptivities, and highly reversible reductive characteristics.
Fabrication of efficient bilayer solar cells with a subphthalocyanine (SubPc) donor supports this
class of materials as promising electron acceptors in organic photovoltaic cells.
As organic materials have low charge carrier mobility, organic and inorganic material combi-
nations have also been used to enhance performance of photovoltaic devices. Salafsky (1999) used
approximately 100 nm thin titanium dioxide nanocrystals with conjugated polymer, poly(p-phenylene
vinylene), where active medium was taken as Al/composite/indium tin oxide. It was observed that the
ratio of inorganic materials affects the overall function of the cell.
Tamura et al. (2014) reported good fill factor and power conversion efficiency with tetrabenzo-
porphyrin (BP), a BP-C60 dyad, and PCBM for the p-, i-, and n-layers, respectively, as compared to
1:1 blend film of BP and PCBM as the i-layer in the p-i-n organic photovoltaic cell. The OPV devices
made of different photoactive layers and the passivated TiO2 electron extraction layers show sig-
nificant enhancement of more than 30% in their power conversion efficiencies. Vasilopoulou et al.
(2014b) prepared TiO2 films through applying alumina (Al2O3) or zirconia (ZrO2) insulating nano-
layers by thermal atomic layer deposition (ALD) and used them as cathode interlayers in organic
solar cells. However, this results in undesirable recombination and a high electron extraction barrier,
thus reducing the open-circuit voltage and the short-circuit current of the complete OPV device.
Lee et al. (2014) fabricated a TiO2-based solar cell using iodine-doped Cu-based metal—organic
frameworks (Cu-MOFs, copper(II) benzene-1,3,5-tricarboxylate) as an active thin layer using a layer-
by-layer technique. Iodine-doped MOFs gave excellent cell performance with Jsc = 1.25 mA cm−2
and Eff = 0.26% under illumination of 1 sun radiation, while the cell with an undoped MOF layer
exhibited only Jsc = 0.05 mA cm−2 and Eff = 0.008%.
Kim et al. (2014b) observed an increase in charge carrier mobility by heteroatom substitution of a
silicon atom by a germanium atom in donor–acceptor-type low-bandgap copolymers, poly[(4,4′-bis(2-
ethylhexyl)dithieno[3,2-b:2′,3′-d]silole)-2,6-diyl-alt-(2,1,3-benzothiadiazole)-4,7-diyl] (PSiBTBT) and
Organic Photovoltaic Cells 67

poly[(4,4′-bis(2-ethylhexyl)dithieno[3,2-b:2′,3′-d]germole)-2,6-diyl-alt-(2,1,3-benzothiadiazole)-4,7-
diyl] (PGeBTBT). Charge carrier mobility was increased because C–Ge bond length is longer in com-
parison to that of C–Si, which modifies the molecular conformation and leads to a more planar chain
conformation in PGeBTBT than in PSiBTBT. This increase in molecular planarity leads to enhanced
crystallinity and an increased preference for a face-on backbone orientation.
The ultraviolet (UV) and ozone treatment increases the power conversion efficiency and air
stability of OPV cells. Le et al. (2014) prepared MoS2 nanosheets by a simple sonication exfoliation
method and used them to form a hole extraction layer (HEL). The OPV cells with MoS2 layers show
a power conversion efficiency of 1.08%, which is less than OPV cells without HEL (1.84%). After
UV/ozone (UVO) treatment of the MoS2 surface for 15 min, the efficiency value increases to 2.44%,
and the work function of MoS2 increases from 4.6 to 4.9 eV. When poly(3,4-ethylenedioxythiophen
e):poly(styrene sulfonate) (PEDOT:PSS) was inserted between MoS2 and the active layer and used a
hole extraction layer, the power conversion efficiency was increased to 2.81%, and air stability also
increased in comparison to the device employing only PEDOT:PSS.
Cathode interfacial material (CIM) was found to be effective in improving the power conversion
efficiency and long-term stability of an organic photovoltaic cell that utilizes a high work function
cathode. Ultraviolet photoemission spectroscopy studies proved that CIM lowers the work func-
tion of the Ag metal as well as ITO and highly oriented pyrolytic graphite (HOPG), and facilitates
electron extraction in OPV devices. Tan et al. (2014) prepared CIM by combining triarylphosphine
oxide and a 1,10-phenanthrolinyl unit, which has high Tg of 116°C and attractive electron trans-
port properties. The characterization of Ag or Al cathodes involving photovoltaic devices shows
improvement in PCE as compared to the reference Ag device and compares well to that of the
Ca/Al device. The CIM/Ag photovoltaic device with the active layer PTB7:PC71BM shows 7.51% of
PCE. The PCE was further increased to 8.56% for the CIM/Al device (with Jsc = 16.81 mA cm−2,
Voc = 0.75 V, and FF = 0.68).
In recent years, graphene has offered a vast range of photonic and electronic applications, such as
photodetectors, optical modulators, high-speed transistors, electromagnetic wave shieldings, notch
filters, linear polarizers, electrochemical energy storage, transparent and flexible electronics, dis-
plays, optoelectronics, sensors, nanoelectromechanical systems, energy technologies, and photovol-
taic cells, which makes it a promising material. Its versatility provides an entirely new generation of
technologies beyond the limits of conventional materials. It has a honeycomb-like structure, which
is made up of a single layer of carbon atoms. Graphene is also attracting the interest of researchers
because it is lightweight, flexible, atomically thin, mechanically strong, electrically tunable, visu-
ally transparent, extraordinarily high in charge mobility and saturation velocity, which helps in
attaining a fast switching speed for radio frequency analog circuits. Untreated graphene is a semi-
metal that does not provide true off state; as a result, it typically precludes its use in digital logic
electronics without bandgap engineering. It becomes highly conducting on doping.
One important characteristic of graphene is that it strongly interacts with light in the microwave
range to the ultraviolet range, with spanning wavelengths of at least five orders of magnitude.
Because of this exception, light interaction between light and graphene, it shows excellent elec-
tronic and mechanical properties. This is why graphene is a promising candidate for many kinds of
photonic devices (Weiss et al. 2012; Xia et al. 2013). Graphene derivatives are in demand because
of their excellent high electronic and thermal conductivity, high specific surface area, and optical
transparency, combined with exceptionally good mechanical flexibility and environmental stabil-
ity. Graphene is also used in photovoltaic, electronic, and electrochemical energy storage (Lee
et al. 2013).
Graphene is a 1 atom–thick layer of graphite with a two-dimensional sp2-hybridized carbon
network. Graphene has an excellent role in improving the overall performance of OPV devices
because of its peculiar properties, such as good mechanical strength, high thermal conductiv-
ity, superior transparency, large specific surface area, and tremendous charge transport properties
(Chang et al. 2014).
68 Solar Energy Conversion and Storage

Materials with a highly ordered phase, such as polythiophene (Too et al. 2001; Brabec et al.
2001b), offer high charge mobility, which increases the performance of the cell. Fluorination of
polythiophene derivatives also affects the performance of organic photovoltaics. Recently, Jo et al.
(2014) reported an increase in the power conversion efficiency by 20%–250% by fluorine atom
substitution in poly(3,4-dialkylterthiophenes) (PDATs). Fluorinated PDATs show a deeper HOMO
energy level than nonfluorinated ones, thus leading to higher open-circuit voltage in organic photo-
voltaic cells and enhanced molecular ordering.
Recently, Takao et al. (2014) used fluorinated subnaphthalocyanine derivatives as donor materi-
als and fullerene as an acceptor for low-molecular-weight organic photovoltaic cells. Such cells
have the low-lying HOMO energy levels, keeping the strong absorption band of long wavelength for
improvement of open-circuit voltage without the expense of short-circuit current density. The fron-
tier orbital energy levels can be effectively tuned by fluorination because the HOMO/LUMO energy
levels of hexafluoro-, heptafluoro-, dodecafluoro-, tridecafluoro-, and the parent subnaphthalocya-
nine were estimated to be 5.69/3.93, 5.67/3.90, 5.96/4.19, 5.92/4.11, and 5.30/3.58 eV, respectively.

5.6  TYPES OF ORGANIC PHOTOVOLTAIC CELLS


5.6.1 Single Layer
Single-layer organic photovoltaic cells (Figure 5.2) are the simplest form, which are made up as
a sandwich model, like a layer of organic electronic materials between two metallic conductors.
Basically, both layers have different work functions, which help to set up an electric field in the
organic layer. Typically, the first electrode is made up of an indium tin oxide layer (ITO) with high
work function, and the second electrode is made up of a metal with low work function such as alu-
minum, magnesium, or calcium. When the organic layer absorbs light, electrons are excited from
HOMO to LUMO, thereby forming excitons. Differences in work functions help to split the exciton
pairs, pulling electrons to the positive electrode (an electrical conductor used to make contact with
a nonmetallic part of a circuit) and holes to the negative electrode.
Ghosh et al. (1974) reported an Al/Mg phthalocyanine/Ag sandwich solar cell model with effi-
ciency of 0.01%. Later, efficiency of 0.30% with a polyacetylene (CH)x -based single-junction solar
cell was reported by Weinberger et al. (1982).
Marks et al. (1994) made the single-layer device structure of OPV cells, which consists of a
transparent electrode/organic photosensitive semiconductor/electrode. They used 50–320  nm
thick poly(p-phenylene vinylene) (PPV) sandwiched between an indium tin oxide (ITO) anode
and a low work function cathode. The quantum efficiency of this cell was reported to be approxi-
mately 0.1% under 0.1mW cm−2 intensity. Charge mobility in an organic semiconductor is around
10−3 cm2 V−1·s−1, which is far less than the mobility of a single-crystalline silicon, which is about
103 cm2 V−1·s−1, resulting in low quantum yield.

Electrode 1
(ITO, metal)

Organic electronic material


(small molecule, polymer)

Electrode 2
(Al, Mg, Ca)
FIGURE 5.2  Structure of single-layer cells.
Organic Photovoltaic Cells 69

Single-layer organic material efficiency was reported as approximately 2%, because of its low
stability of material, poor charge transport, and limitation by the low red light absorption. New
materials were synthesized to solve these problems by use of material combination, optimization
of molecular design, or self-assembly processing in order to control morphology, which increased
performance up to 5% (Nelson 2002).
The morphology or architecture of semiconductor blends shows vast impact on the performance
of the cells. It means cells should be prepared in such a way that they can capture more sunlight in
order to produce more electricity or photocurrent (McGehee and Topinka 2006). Electron transfer
along with effective hole transfer increase efficiency by enhancing short-circuit current. Wang et al.
(2014b) showed that organic photovoltaic cells based on ambipolar donor–acceptor1–acceptor2
architectural 4-styryltriphenyl amine exhibit a long-lived charge separation state with a lifetime
of 650 ns. Not only the morphology but also the conditions of preparation of organic film affect
the current–voltage characteristics in a positive way. When organic film was prepared with copper
phthalocyanine (CuPc) and hexadecafluoro CuPc (F16CuPc) under different conditions, they then
show different efficiencies of cells.
Page et al. (2014) reported that the thickness of the interlayer does not affect the performance
of the single-junction solar cell. They prepared a single-junction solar cell based on the fulleropyr-
rolidines with amine (C60-N) or zwitterionic (C60-SB) substituents as cathode-independent buffer
layers ranging from 5 to 55 nm thickness. It was observed that the effective work function of Ag,
Cu, and Au electrodes was reduced by using a thin layer of C60-N to 3.65 eV, and more than 8.5%
efficiency was obtained, but this result was independent of cathode material (i.e., Al, Ag, Cu, or Au).
Beliatis et al. (2014) used PCDTBT:PC70BM with solution-processed nanostructures of metal
oxide–reduced graphene oxide (RGO) as electron transport layers (ETLs) exhibited the efficiency
of 8% in single cells. The power conversion efficiencies were reported to be about 8% and 9%,
respectively, using polymer and small molecules to synthesize a single-junction organic solar cell.
Zhang et al. (2015) reported organic photovoltaic devices with oligothiophene-like small molecules,
consisting of seven conjugation units as the backbone and 2-(1,1-dicyanomethylene)rhodanine as
the terminal unit (i.e., DRCN7T) and [6,6]-phenyl C71-butyric acid methyl ester (PC71BM) as the
acceptor unit. The photocurrent generation efficiency was observed as 9.30%. The DRCN7T-based
cells showed exceptionally high internal quantum efficiency (100%). That may be because of a
nanoscale donor–acceptor network or highly crystalline donor fibrils with approximately 10  nm
diameters, which is close to diffusion length of the organic material and an efficient electron trans-
port layer. Hybrid graphene–metal oxide materials are used as improved ETLs to improve effi-
ciency by enhancing the charge transport process.
In a single-layer OPV cell, charge dissociates into free carriers only at one place (i.e., in the interface
between semiconducting organics and a cathode). Later, it was investigated that the excitons are more
efficiently dissociated at the interface between donor and acceptor, so a bilayer OPV was developed.
Single-layer organic photovoltaic cells have low quantum efficiency of less than 1% and low PCE
of less than 0.1%. The major drawback with these cells is that the electric field generated from the
difference between the two conductive materials is rarely enough to split an exciton. Thus, it some-
times results in electron–hole recombination.

5.6.2 Bilayer
Bilayer cells contain two layers with different electron affinities and ionization energies in between
the conductive electrodes, which generate electrostatic forces at the interface between the two lay-
ers. The layer with higher electron affinity and ionization potential is the electron acceptor; the other
layer is the electron donor. This structure is also called a planar donor–acceptor heterojunction. A
bilayer OPV cell can be prepared by inserting an acceptor layer between a donor semiconducting
organic and a cathode. It means the bilayer organic photovoltaic cell has an additional electron-
transporting layer, which is not found in the single-layer OPV structure (Figure 5.3).
70 Solar Energy Conversion and Storage

Electrode 1
(ITO, metal)

Electron donor

Electron acceptor

Electrode 2
(Al, Mg, Ca)
FIGURE 5.3  Structure of bilayer cells.

The interface between the two organic materials plays major role in determining the properties
of the cell electrode/organic interlayer. The thin-film bilayer OPV cell structure was first realized
by Tang (1986). The device consists of indium tin oxide (ITO)/copper phthalocyanine (CuPc)/per-
ylene tetracarboxylic derivative (PV)/silver (Ag) with a power conversion efficiency of 1% under
simulated AM2 conditions. This 10-fold increase in PCE resulted from improving exciton dissocia-
tion efficiency by adding an electron-transporting material that forms an offset energy band with
the hole-transporting material. Antohe and Tugulea (1991) prepared a two-layer organic cell from
copper phthalocyanine as a p-type organic semiconductor and 5,10,15,20-tetra(4-pyrydil)21H,23H-
porphyne (TPyP) as an n-type organic semiconductor.
Sariciftci et al. (1993) prepared the bilayer organic photovoltaic cell for the first time using a
conjugated polymer, where poly[2-methoxy-5-(2′-ethyl-hexyloxy)-1,4-phenylene vinylene] MEH-
PPV absorbs light and transports holes to the anode, and fullerene acts as the electron-transporting
material to the cathode. A power conversion efficiency of 0.04% was reported under monochromatic
incident light at 514.5 nm, which was slightly higher than single polymer layer PV cells, but the
performance is still low due to the intrinsically short diffusion length of excitons in organic semi-
conductors (Halls et al. 1996; Roman et al. 1999; Theander et al. 2000). To overcome this problem,
Peumans et al. (2003) replaced perylene tetracarboxylic derivative with C60 as an acceptor in the
device structure, which increased the device efficiency to 3.5%.
Zimmerman et al. (2013a) reported that replacing the exciton-quenching buffer layer by an exci-
ton-blocking layer enhances the performance of the bilayer solar cell. When an exciton-blocking
benzylphosphonic acid (BPA)–treated MoO3 or NiO layer was used in place of the exciton-quenching
MoO3 anode buffer layer in bilayer organic photovoltaic cells, then the power conversion efficiency
shows significant improvement. The addition of untreated MoO3 anode buffers and BPA-treated
NiO buffers in diphenylanilo-functionalized squaraine (DPSQ)/C60-based bilayer devices show effi-
ciency from 4.8% ± 0.2% to 5.4% ± 0.3% under illumination of 1 sun AM1.5G. Further addition
of a highly conductive exciton-blocking bathophenanthroline (BPhen):C60 cathode buffer increases
cell performance up to 5.9% ± 0.3%.
Use of fullerene as an electron donor increases the performance of planer heterojunction (PHJ)
organic photovoltaic cells (Zhuang et al. 2013). Organic photovoltaic cells were prepared with elec-
tron donor, fullerene derivatives indene-C60 bisadduct (ICBA), and phenyl C61-butyric acid methyl
ester with fullerene C70 as the electron acceptor. Two processes are included: (1) fullerene bulk
includes charge generation and (2) exciton dissociation at the donor–acceptor interface. Indene-C60
bisadduct with 5 nm thickness and C70 fullerene with 40 nm thickness give external quantum effi-
ciency on an exposing long wavelength photon. This means total efficiency depends on the thick-
ness of the donor fullerene.
A double layer of transparent conducting oxide (TCO) films increases the power conversion effi-
ciency as compared to a single layer (Cho et al. 2014). It has been observed that indium tin oxide (ITO)
and aluminum-doped zinc oxide (AZO) single-layered films are less efficient than the ITO/AZO films
Organic Photovoltaic Cells 71

of 500/250 nm thickness, while the double layer exhibits higher photocurrent due to higher transmit-
tance and lower resistance.
A short rapid thermal annealing of methylammonium lead mixed halide perovskite (CH3NH3PbI3–xClx)
at 130°C leads to the growth of large micron-sized textured perovskite domains. It also increases the
short-circuit currents and power conversion efficiencies to 13.5% for the planar heterojunction perovskite
solar cells (Saliba et al. 2014). Simultaneous optimization of light absorption and carrier collection
increases the short-circuit current in thin planar organic photovoltaic heterojunction cells. Tsai et al.
(2014) observed that an ultrathin-film solar cell with a SubPc/C60 photovoltaic structure shows 78%
power conversion efficiency and short-circuit current of 0.790 mA cm−2 for a 30 nm thick cell, but only
32% power conversion efficiency for a 45 nm thick cell, for which short-circuit current is 0.980 mA cm−2.
Typically, a diffusion length of organic material is 10 nm. For exciton diffusion to the interface
of the layer and for generation of the charge carrier, the thickness of the layer must be equal to or
in the same range as the diffusion length. However, polymer layer thickness should be 100 nm to
absorb a sufficient amount of light, and at such a thickness, only a small amount of the excitons can
reach the heterojunction interface.

5.6.3 Heterojunctions
Bulk heterojunction (BHJ) is made up of a nanoscale blend of donor and acceptor layers (Figure 5.4).
These cells are thick enough to absorb a broad range of spectra. Yu and Heeger (1995) made a
phase-separated polymer blend of donor poly[2-methoxy-5-(2′-ethyl-hexyloxy)-l,4-phenylene vin-
ylene] (MEH-PPV) and acceptor cyano-PPV (CN-PPV). Scharber et al. (2006) prepared bulk het-
erojunction solar cells from conjugated polymers and a fullerene derivative. They also exhibited
the relation between the open-circuit voltage and the oxidation potential for different conjugated
polymers.
Li et al. (2005) showed significant enhancement in power conversion of a blend of OPV cells
by using poly(3-hexylthiophene) (P3HT) as a hole-transporting polymer and PCBM as a soluble
C60 derivative, while Kim et al. (2006) prepared a blend of organic photovoltaic cells using various
P3HT and PCBM blend solutions. The P3HT solution had different regioregularities ranging from
80% to 96%. The P3HT solution with a higher order of regioregularity resulted in the crystallized
fibril-like shape, which increased charge transport as well as photon absorption efficiency.
Chasteen et al. (2008) prepared photovoltaic cells consisting of layers and blends of a hole-
transporting derivative of poly(p-phenylene vinylene) with different electron transporters, such as
titanium dioxide, a cyano-substituted PPV, and a fullerene derivative (PCBM), to increase the per-
formance of device. They studied time-resolved and steady-state photoluminescence phenomena
and found that morphological differences such as chain conformation or domain size, often over-
shadowed the effect of charge transfer and also decreased losses to recombination. The charge
transfer and excitons dissociation was found to increase by the addition of PCBM. But the electron-
transporting polymer CN-ether-PPV does not show the same results. On increasing CN-ether-PPV

Electrode 1
(ITO, metal)

Dispersed heterojunction

Electrode 2
(Al, Mg, Ca)
FIGURE 5.4  Structure of heterojunction cells.
72 Solar Energy Conversion and Storage

in the polymer/polymer blend, large domains were created, which adversely affected the efficiency
of the device. Therefore, a layered device structure is more efficient than a blended one because
morphology is more easily controlled in a layered device.
The bilayer OPV cell collects a very small amount of excitons, which are created near the inter-
face of the donor and the acceptor, leading to the development of bulk heterojunction OPV cells.
In such cells, there is an intermixed composite of donors and acceptors that have a larger interface
area. Ko et al. (2009) reported an enhancement in the absorption of light for OPV cells by patterning
it using a photonic crystal nanostructure embossed in the photoactive bulk heterojunction layer via
PRINT. This results in not only greater adsorption but also electrical enhancements. This method
lends itself to a 4 cm2–area fabrication of nanoscale features. The efficiency increased to 70%.
The efficiency of OPV devices depends on molecular ordering and lowering the HOMO level;
this can be accomplished by using the proper processing method. Kim et al. (2013) synthesized
conjugated polymers by the Stille polymerization reaction, in which a symmetrically branched
alkyl side chain benzotriazole (DTBTz) is taken as an acceptor semiconductor, and unsubstituted
or (triisopropylsilyl) ethynyl (TIPS)-substituted 2,6-bis(trimethylstannyl)benzo[1,2-b:4.5-b′] dithio-
phene (BDT) is taken as a donor semiconductor. The optical bandgap was reported to be 1.97 and
1.95  eV for PBDT-DTBTz and PTIPSBDT-DTBTz, respectively. The TIPS groups in the donor
unit is responsible for the molecular ordering and lowering of the HOMO level, which improves
the efficiency of the solar cell. Bulk heterojunction photovoltaic cells prepared from conjugative
polymers (triisopropylsilyl) ethynyl (TIPS)-substituted 2,6-bis(trimethylstannyl)benzo[1,2-b:4.5-b′]
dithiophene (BDT) and symmetrically branched alkyl side chain benzotriazole with power conver-
sion efficiency of 5.5%, whereas 2,6-bis(trimethylstannyl)benzo[1,2-b:4.5-b′] dithiophene (BDT)
and benzotriazole with symmetrically branched alkyl side chains show a power conversion effi-
ciency of only 2.9%. This means that the TIPS group is responsible for higher optimal morphology
and carrier mobility in OPV cells.
A bulk heterojunction is the most active layer system for and OPV cell. A BHJ is a network
of domains that contains blended donor and acceptor molecules, often in pure and mixed phases
(Pfannmoller et al. 2013). A bulk heterojunction with two-component p- and n-doped materials
exhibits power conversion efficiencies in excess of 7% (Chen et al. 2013).
The interface structure is directly related to the performance of the cell, as it affects the recom-
bination rate for electrons and holes at donor–acceptor heterojunctions in thin-film OPV cells. Cells
must have disorder at the heterointerface and order in the bulk of the thin films (epitaxial rela-
tionships) to achieve high short-circuit current and open-circuit voltage. A squaraine donor and
C60 acceptor heterojunction cell enhances the charge recombination through epitaxial relationships
and provides interdiffusion, which reduces open-circuit voltage and improves short-circuit current
(Zimmerman et al. 2013b).
Hedley et al. (2013) observed that morphology of bulk heterojunction plays an important role in
the performance of OPV devices. A PTB7:PC 71 BM blend is composed of fiber-like, elongated,
fullerene-rich and polymer-rich domains, which are 200–400 nm long and 10–50 nm wide. This
morphology provides an efficiency of 80%, as it allows a concentration gradient for directional
charge diffusion that leads to the extraction of charge pairs. But a decrease in efficiency to 45% was
observed when agglomerated fullerene was used with a blend instead of elongated fullerene.
Yu and Chan (2013) prepared bulk heterojunction hybrid photovoltaic cells by incorporating a
p-type NiO thin layer in between indium tin oxide (ITO)/nickel oxide (NiO)/poly(3-hexylthiophene)
(P3HT): [6,6]-phenyl C61-butyric(PCBM):titania (TiO2): platinum (Pt) nanoparticles (NPs)/Ca/Al
layers. It was observed that the NiO interlayer enhances stability and performance of the cell. The
optimum cell performance of ITO/NiO of 5 nm thickness and P3HT:PCBM:TiO2 of 15 wt.% : Pt
of 0.03 wt.% or Ca /Al was recorded as 2.1% with an open-circuit voltage of 0.61 V, short-circuit
current density of 6.22 mA cm−2, and fill factor of 54.8%. Paci et al. (2013) incorporated poly(3-
hexylthiophene) nanofibers in the polymer–fullerene nanostructured films to improve the structural
durability.
Organic Photovoltaic Cells 73

Two electron-withdrawing fused thiadiazole units, made by fusing quinoxaline or phenazine with
a benzothiadiazole unit, thiadiazolo[3,4-g]quinoxaline (DTBTQx), and thiadiazolo[3,4-i]phenazine
(DTBTBPz) and a dithiophene electron-donating unit, were developed by Stille reaction. Two copo-
lymers in the film state exhibited broad absorption spectra with a narrow bandgap of approximately
1.2 eV. A bulk heterojunction solar cell shows a power conversion efficiency of 0.52%, with a Jsc
of 1.44 mA cm−2, Voc of 0.69 V, and FF of 0.40, after thermal annealing at 90°C (Hai et al. 2014).
Owczarczyk et al. (2014) carried out synthesis of push–pull conjugated copolymers based on
cyclopenta[c]thiophene-4,6-dione (CTD) and benzodithiophene (BDT) as an organic photovolta-
ics electron donor material. Conjugated polymers have been used as electron donors and fullerene
derivatives as electron acceptors. The power conversion efficiency of such a polymer solar cell has
now exceeded 10%. Quinoidal units, phospholes, porphyrins, and fluorinated aromatic rings have
been used as building blocks, which can now be introduced into low-bandgap conjugated polymers
(Umeyama and Imahori 2014).
Shivanna et al. (2014) made a nonfullerene-based bulk heterojunction organic solar cell using
a nonplanar perylene dimer (TP) as an electron acceptor and a thiophene-based donor polymer
poly{[4,8-bis-(2-ethyl-hexyl-thiophene-5-yl)-benzo[1,2-6:4,5-6′] dithiophene-2,6-diyl]-alt-[2-(2′-
ethyl-hexanayl)-thieno[3,4-6] thiophen-4, 6-diyl]} (PBDTTT-CT). The mixture of the donor poly-
mer and TP at 50:50% weight ratio showed a photon-to-current conversion efficiency of 45% in the
visible region and a power conversion efficiency of 3.2%.
Chan et al. (2014) synthesized a new p-type perylene diimide as a photoactive material in
OPV devices. This nonplanar and three-dimensional spirobifluorene-modified perylene diimide
compound can be used as a donor material in combination with fullerene to form bulk hetero-
junctions. This new photoactive material shows high open-circuit voltages of 0.97 V and a power
conversion efficiency of up to 4%. The combination of isoindigo as the electron-deficient acceptor
and 3,4-ethylenedioxythiophene as the electron-rich donor, followed by CH-arylation with different
acceptors (4,7-dibromo[c][1,2,5]-(oxa, thia, and/or selena)diazole, were used to develop low-bandgap
donor–acceptor–donor–acceptor (D-A-D-A) polymers. These polymers have high stability and
good solubility in chlorinated solvents. Elsawy et al. (2014) utilized poly ((E)-6-(7-(benzo-[c]
[1,2,5]-thiadiazol-4-yl)-2,3-dihydrothieno-[3,4-b][1,4]dioxin-5-yl)-6′-(2,3-dihydrothieno-[3,4-b]
[1,4]-dioxin-5-yl)-1,1′-bis-(2-octyldodecyl)-[3,3′-biindolinylidene]-2,2′-dione) as the donor and
PC61BM as the acceptor in BHJ with a short-circuit current density of 8.10 mA cm−2, open-circuit
voltage of 0.56 V, fill factor of 35%, and PCE of 1.6%. These polymers were used as donor materi-
als for photovoltaic applications, mainly in polymer solar cells.
Solution-processed photovoltaic cells synthesized with electron donor material 3,6-bis(N,N-
dianisylamino)-fluoren-9-ylidene malononitrile (FMBDAA36) and ITO/PEDOT:PSS/ (1:3[w/w]
FMBDAA36:PC71BM)/LiF/Al show performance to 4.1%, open-circuit voltage of 0.89  V, short-
circuit current of 10.35 mA cm−2, and fill factor of 44.8%. Lim et al. (2014) studied the impact of
ultrathin interfacial metal fluoride in the performance of bulk heterojunction organic photovoltaic
cells. They used an ultrathin BaF2 single-coverage layer (less than 3 nm) at the electron extraction
contact, which showed a positive effect on the open-circuit voltage and power conversion efficiency
of the organic photovoltaic cells, but the short-circuit current remained almost constant. An effi-
ciency of 4.0% was obtained in the presence of the interlayer, but efficiency of only 2.1% was shown
in its absence (under 100 mW cm−2 solar radiation). The enhancement in performance may be due to
a hugely improved lifetime and lowered effective work function of the cathode caused by the large
dipole moment of thin BaF2 films.
Solvent-processed bulk heterojunction solar cells composed of 2,4-bis[4′-(N,N-di(4′-hydroxyphenyl)
amino)-2′,6′-dihydroxyphenyl]squaraine (Sq-TAA-OH) as an active layer with configuration ITO/
PEDOT:PSS/Sq-TAA-OH:PC71BM/LiF/Al were reported (Karak et al. 2014). The bandgap and
HOMO level of Sq-TAA-OH were reported to be 1.4 and –5.3 eV, respectively. The power conver-
sion efficiency of about 4.8% was observed by a reproducible procedure using a mixture of good
and poor Sq-TAA-OH-solubilizing organic solvents along with a bulk heterojunction layer made up
74 Solar Energy Conversion and Storage

of diiodooctane (DIO), followed by thermal annealing. The DIO is a good and poor solvent mixture
that broadens the Sq-TAA-OH crystallites’ size distribution in pristine films, whereas narrower size
distribution is provided by thermal annealing.
Kim et al. (2014c) used difluorinated benzoselenadiazole (DFDTBSe) and ethylhexyloxy (EH)-
or octyldodecyloxy (OD)-substituted benzo[1,2-b:4,5-b′]dithiophene (BDT) as monomer units to
form copolymers. These conjugated copolymers have low bandgaps of 1.66 and 1.69 eV and HOMO
energy levels of –5.44 and –5.43 eV, respectively. The molecular packing and order of the active
layer blended with [6,6]-phenyl-C71 butyric acid methyl ester (PC71BM) depend on the different
alkyloxy side chain in the polymer. The PODBDT-DFDTBSe:PC71BM film is composed of an
edge-on structure, while the PEHBDT-DFDTBSe:PC71BM film has a face-on structure. It was also
reported that a bulk heterojunction cell fabricated with a PEHBDT-DFDTBSe:PC71BM active layer
exhibited a maximum power conversion efficiency of 5.74%, which is the highest efficiency among
benzoselenadiazole and BDT-derivative polymeric materials containing OPCs.
In bulk heterojunction OPV cells, the benzoselenophene derivative electron donor with PC61BM
as the electron acceptor shows a power conversion efficiency as high as 5.8% (Park et al. 2014).
Stanculescu et al. (2014) prepared a heterojunction structure using single layers of arylene-based
polymer, poly[N-(2-ethylhexyl)2.7-carbazolyl vinylene]/AMC16 and poly[N-(2-ethylhexyl)2.7-car-
bazolyl 1.4-phenylene ethynylene]/AMC22 with Buckminsterfullerene/C60 in the weight ratio of 1:2
(AMC16:C60) and 1:3 (AMC22:C60) by the matrix-assisted pulsed laser evaporation (MAPLE) tech-
nique. Different heterojunctions were prepared on glass/ITO/PEDOT-PSS with AMC16, AMC22,
and AMC22:C60 layer by MAPLE. The highest performance was shown by a single layer of AMC16
polymer: glass/ITO/PEDOT-PSS/AMC16/Al.
The combination of a high fullerene-C70 and fullerene-C60 as amorphous donor and acceptor,
respectively, in an optimized tandem configuration results in efficiency of 7.2%. On the contrary,
a single heterojunction cell with donor–acceptor layers exhibits efficiency above 5.5%, with its
material of 1.8–1.4 eV bandgap. It shows that the fill factor of the tandem stack is higher in com-
parison to either one of the subcells (Cheyns et al. 2014). In bulk heterojunction OPV cells, trans-
parent conducting electrodes indium tin oxide (ITO) can be replaced by multilayer transparent
electrode MTO/Ag/MTO (MAM) with a nanosized Ag thin film embedded between Mn-doped tin
oxide (MTO) layers. It was prepared by employing a RF sputtering process at room temperature.
The MTO/Ag/MTO multilayer electrodes are more advanced than an indium tin oxide (ITO) elec-
trode as it exhibits transmittance of 80.1%–85.4% within a 380–780 nm wavelength range, a sheet
resistance of 10.1–10.6  Ω/sq, a short-circuit current of 7.12  mA  cm−2, an open-circuit voltage of
0.62 V, a fill factor of 0.62, and a power conversion efficiency of 2.73% (Lee et al. 2015).
Aziz et al. (2015) used green color dye vanadyl 2,9,16,23-tetraphenoxy-29H,31H-phthalocyanine
(VOPcPhO) in between a photoactive layer of VOPcPhO, which has been sandwiched between
indium tin oxide (ITO) and aluminum (Al) to fabricate ITO/PEDOT: PSS/VOPcPhO/Al photovol-
taic devices. The device showed a photovoltaic effect with Jsc of 5.26 × 10−6 A cm−2, Voc of 0.621 V,
and FF equal to 0.33. It had a limitation in that when a single layer of vanadyl phthalocyanine
derivative was incorporated in the solar cells, the efficiency decreases. On the contrary, adding a
variety of donor materials to form bulk heterojunction increases the efficiency of the cell.
The photoactive layer made up of homogeneously dispersed carbon nanotubes (CNTs) in a semi-
conducting polymer–fullerene derivative bulk heterojunction matrix shows improvement in optical
and electrical properties due to effects of the wide-band photoabsorption and high charge carrier
mobility of the CNTs. The CNTs are functionalized by alkyl amide groups for high dispersion in
organic media and by homogeneously mixing them with the polymer and fullerene in solution. This
exhibits a 40% (3.2%–4.4%) increase in power conversion efficiency as compared to an organic
solar cell using a photoactive layer without CNTs. The open-circuit voltage decreases slightly due
to the small energy level variation of the active layer, when CNTs were used in the semiconducting
polymer poly 3-hexylthiophene and a fullerene derivative, [6,6]-phenyl C61 butyric acid methyl ester
(P3HT:PCBM) as photoactive materials. It was concluded that power conversion efficiency increases
Organic Photovoltaic Cells 75

due to an increase in short-circuit current from 10.5 to 14.6 mA cm−2. The photocurrent enhance-
ment in the CNT-incorporated photoactive layer is attributed to two major causes: (1) an increase in
the absorbed photons, and (2) an increase in the charge carrier mobility (Gwang et al. 2012).
The single-walled carbon nanotube (SWNT) also affects the efficiency of the cell. It may decrease
or increase the performance depending on the length of the single-walled carbon nanotube. It has been
reported that when full-length SWNTs were added between poly(3-hexylthiophene)-[6,6]-phenyl-
C61-butyric acid methyl ester (P3HT-PCBM) (1:1 w/w) active layers in the bulk heterojunction cell, then
efficiency decreases (only 1.3%) at 80°C –225°C. But a cell without SWNTs shows efficiency to 2%.
When shortened SWNTs were incorporated, it was noticed that efficiency reached 2% and increased
by nearly 50% at 70°C. This enhancement is due to improved hole transport through the SWNTs in
devices (Skupov and Adronov 2014).
The morphology or structure showed a major influence on the photocurrent power of organic
photovoltaic devices. The heterojunction cell made up of zinc phthalocyanine/fullerene (ZnPc/C60)
with poly(3,4-ethylenedioxythiophene):poly(styrenesulfonate) and 2,5-bis(4-biphenylyl)-bithiophene
(BP2T) between indium tin oxide and ZnPc layers showed much better efficiency than a conventional
cell without BP2T (i.e., almost 120% higher efficiency). With the addition of BP2T in ZnPc/C60
bilayer films, a power current efficiency of 2.6% was obtained as a result of improved short-circuit
current and fill factor. It was also noticed that when ZnPc films were tuned with long fiber-like grains,
they show more photocurrent generation because the charge carriers can transport for a longer dis-
tance in comparison to the round and short fiber-like grains (Wang et al. 2014c).
Recently, new cascade heterojunction (CHJ) organic solar cells have been investigated, which
are the best replacement for series-connected tandems and conventional bulk heterojunctions. Such
cells have poor fill factor and minimal enhancements in power conversion efficiency, but show
high internal quantum efficiency (IQE) and broad spectral coverage. The fill factor and efficiency
of the cells can be increased by tuning the maximum power point voltage of the constituent par-
allel-connected heterojunctions and reducing the intrinsic injection barriers. It was recorded that
a CHJ device with a transparent exciton dissociation layer (EDL)/interlayer/acceptor architecture
with more than 99% internal quantum efficiency in the interlayer showed a 46% increase in power
conversion efficiency as compared to a single heterojunction (SHJ).

5.7  TANDEM SOLAR CELLS


Tandem solar cells consist of two staked single subcells, which are connected by an interlayer.
Basically, a tandem solar cell provides a new way to capture or harvest a broad range of solar radia-
tion spectrum, as it involves combining two or more solar cells with different absorption bands.
Generally, a polymer single-layer cell has less efficiency because of the lack of a suitable low-
bandgap polymer. Therefore, polymer tandem solar cells provide an innovative way to achieve high
efficiency. The low-bandgap polymer decreases the performance of the cell, but a combination
of such polymers shows an increase in the efficiency of the cell reported. The performance of
an organic tandem cell can be increased (Hadipour et al. 2006) by electronically combining two
bulk heterojunction single subcells in series to prepare a solution-processed polymer tandem cell.
Two heterojunctions have maximum absorption near 850 nm and 550 nm in this cell. The cell was
designed in such a way that the thickness of the bottom cell layer was tuned to maximize the optical
absorption of the top cell layer. The combination concept results in a high open-circuit voltage (i.e.,
equal to the sum of each subcell [1.4 V]).
Tvingstedt et al. (2007) reported a doubling of the power conversion efficiency of 1.8 ± 0.3 by
folding two conjugative polymer planar cells, which were spectrally different. The tandem organic
solar cell performance also depends on the interlayer and illuminated light. Zhao et al. (2008)
observed that a polymer–small molecule tandem cell using high transparent Al and a MoO3 inter-
mediate layer of 1 nm Al and 15 nm MoO3 thickness exhibited efficiency of 2.82% on illumination
under 100 mW cm−2, a short-circuit current density of 6.05 mA cm−2, and the open-circuit voltages
76 Solar Energy Conversion and Storage

of individual cells of 1.01 V. The power conversion efficiency increases to 3.88% when the tandem
cell was illuminated under 300 mW cm−2.
Hadipour et al. (2008) studied a multilayer tandem solar cell prepared by small molecules and
polymers with their specific absorption maxima and width. The main advantage of such cells is
that the voltage at which charges are collected in each subcell is closer to the energy of the photons
absorbed (i.e., the photon energy is utilized more efficiently). The recombination of small- and
large-bandgap polymer semiconductor cells showed efficiency of 4.9%, and the short-circuit current
was observed to be far greater than the current limiting the subcell (Gilot et al. 2010).
The interlayer is a metal/semiconductor contact, which is different than the traditional tunnel
junction in inorganic tandem cells. This interlayer plays an important role in deciding the per-
formance of such devices. An interlayer of p-n type between two subcells shows a power conver-
sion efficiency of 5.8% (Sista et al. 2010). A robust interlayer is physically quite strong, optically
transparent, and electrically conductive in nature. Yang et al. (2011) reported that on exposing such
cells to light, charges are collected and recombined in the interconnecting layer, while charges are
generated and extracted from the tandem solar cell under dark conditions. Such an interconnecting
layer cell shows high PCE (7.0%). Short-circuit current is affected by a variation in the thickness of
the interlayer, which is the space between two subcells. The variations in thickness of transparent
p-doped interlayer in tandem solar cells results in an enhancement in the short-circuit current but
does not affect (or negligibly affects) open-circuit voltage and fill factor (Riede et al. 2011).
Graphene oxide (GO) is a two-dimensional, random copolymer with two units with distributed
nanosize graphitic patches and highly oxidized domains. Tung et al. (2011) prepared a sticky inter-
layer by casting a mixture of GO and conducting polymer poly(3,4-ethylenedioxythiophene):poly
(styrenesulfonate) (PEDOT:PSS) in water. The insulating GO makes PEDOT highly conductive.
The combination of a tandem solar cell and two vacuum-processed single heterojunctions was
tried by Riede et al. (2011). The main concept behind this combination was to make a tandem
solar cell, which is sensitive to the entire visible range of light. Two heterojunctions were used for
this purpose. The first was C60 and a dicyanovinyl-capped sexithiophene derivative (DCV6T) that
absorbs mainly in the green region, whereas the second incorporates C60 and a fluorinated zinc
phthalocyanine derivative (F4-ZnPc) that absorbs mainly in the red region. Both of these together
make a tandem solar cell, which is sensitive to the entire visible range of the solar spectrum.
Wang et al. (2011) investigated multilayer junction solar cells based on colloidal quantum dots
(CQDs). The electronic bandgap was tuned by combining colloidal quantum dots to make it active
toward a broad range of the solar spectrum. First colloidal quantum dots (CQDs) PbS, tandem solar cell
was also prepared by Wang et al. in which graded recombination was used that permits electron and
hole currents to meet and recombine. The open-circuit voltage equal to the sum of the two constituent
single-junction devices (i.e., 1.06 V) and power conversion efficiency of about 4.2% were obtained.
Dou et al. (2012) made such a call by coupling low-bandgap conjugated polymer PBDTT and
DPP with bandgap of approximately 1.44 eV. A power conversion efficiency was reported of about
8.62%, which was far better than that for a single-layer cell with the same polymer (near 6%).
Li et al. (2013) made tandem OPV cells with a solution-processed intermediate layer (IML)
consisting of ZnO and neutralized poly(3,4-ethylenedioxythiophene):poly(styrenesulfonate)
(PEDOT:PSS) combined in series to obtain excellent functionality, reliability, and performance.
The tandem solar cells containing a processed low-temperature solution and a chemically stable
interconnecting layer such as PEDOT:PSS/AZO/PEIE were far better than single-junction solar
cells, because they can double the open-circuit voltage and utilize photon energy efficiently (Mitul
et al. 2015).

5.8  HYBRID TANDEM PHOTOVOLTAIC CELL


A new generation of photovoltaic solar cells has evolved as a promising and renewable energy source
because of their properties, such as low cost and excellence in optical and electrical properties. In the
Organic Photovoltaic Cells 77

third generation of photovoltaics, semitransparent polymer material was used in hybrid tandem pho-
tovoltaic cells, and a power conversion efficiency of 5.0% was obtained by Beiley et al. (2013).
Jin et al. (2014) synthesized hybrid solar cells with Cd0.75Hg 0.25 Te colloid quantum dots (CQDs)
and water-soluble conjugative polymer by aqueous process. This aqueous-processed solar cell
exhibited excellent sensitivity toward light because a composite of these materials provides an effec-
tive electron transfer and a power conversion efficiency of 2.7% under 100 mW cm−2 illumination. It
showed efficiency as high as 15.04% under the Herschel infrared region (i.e., 780–1100 nm), which
is reported as the highest among organic–inorganic hybrid solar cells (HSCs) made up of Cd xHg1–xTe
CQDs and the highest near-infrared (NIR) contribution among aqueous-processed HSCs.

5.9  INVERTED TANDEM SOLAR CELL


It is known that tandem solar cells are a combination of two or more bulk heterojunctions, active
for different adsorption bands prepared by various processes. This makes the cell unable to use
a wider range of solar absorption capacities. Kim et al. (2007) utilized two bulk heterojunctions
made up of semiconducting polymers and fullerene derivatives by solution processes. These are
separated and connected by a transparent titanium oxide (TiOx) layer. This TiOx layer helps in the
electron transportation and collection on the first layer and provides a stable foundation that forms
a second cell. Efficiency was recorded at more than 6% at illuminations of 200 mW cm−2, when
the front cell consists of the low-bandgap polymer–fullerene composite as the charge separating
layer, and the back cell is composed of the high-bandgap polymer. This architecture is known as
an inverted structure.
A metal-oxide-only MoO3/Al/ZnO interlayer between two polymer bulk heterojunction sub-
cells was used to make inverted-polymer tandem solar cells, which leads to a power conver-
sion efficiency of 5.1% (Chou et al. 2011). The active layer poly(3-hexylthiophene) (P3HT) and
1-(3-methoxycarbonyl)-propyl-1-phenyl-(6,6)C61 (PCBM) and an intermediate layer MoO3/Ag/Al/
Ca were used to prepare an inverted tandem cell (Zhao et al. 2011). The indium tin oxide (ITO)
modified by Ca was used as the cathode (electron collection) for the top subcell and Ca that form
a continuous layer of 3 nm thickness, while Ag was used as the anode for hole collection. Then
MoO3 was used as the anode buffer layer, which helps to reduce exciton quenching and charge loss.
A  2.89% power conversion efficiency was observed under 100  mW  cm−2 under AM  1.5  G solar
irradiation. The short-circuit current density of 4.19 mA cm−2, open-circuit voltage of 1.17 V, and
fill factor of 59.0% were recorded.

5.10  INVERTED ORGANIC PHOTOVOLTAIC CELLS


In spite of the high power conversion efficiency of OPV devices, there are still some limitations in
terms of device stability due to the need for air-sensitive low work function metal cathodes, such
as Al. The diffusion of oxygen into the organic active layer through pinholes and grain boundaries
in the Al cathode may lead to the degradation of the active layer, which results in device instability
in air. One approach to solving this issue is to adopt an inverted structure, where the charge separa-
tion and collection nature of the electrode are reversed.
A poly(2,7-carbazole) derivative (PCDTBT):fullerene derivative (PC70BM) active layer was used
for fabrication of highly efficient and air-stable inverted-polymer photovoltaic cells with molybde-
num oxide and zinc oxide films as hole- and electron-transporting layers, respectively. It showed
the highest power conversion efficiency of 6.3% (Sun et al. 2011). The solution-processed (sol-gel)
vanadium oxides (VOx) act as a hole-transporting and protecting layer in inverted OPV. The vana-
dium oxides (VOx), an interfacial layer, lead to enhancement in performance (Chen et al. 2011). ZnO
thin films on indium tin oxide as electron-collecting layers were fabricated along with ultrathin
TiO2 layers by atomic layer deposition. It was observed that as the thickness of TiO2 increases,
efficiency was found to decrease. The TiO2 layer helps in preventing the electron collection on ZnO.
78 Solar Energy Conversion and Storage

The electron and hole recombination on the surface of ZnO can be quenched by TiO2, resulting in
improved efficiency of the inverted organic solar cells (IOSCs) (Seo et al. 2011).
Interface treatment with 3-aminopropyltriethoxysilane (APTES) leads to an increase in power
conversion efficiency of approximately 8.08%  ±  0.12% with histogram skewness of –0.291 in
PTB7:PC71BM inverted organic photovoltaic cells. On the contrary, an untreated device shows an
efficiency of 7.80% ± 0.26% with histogram skewness of –1.86. A 3-aminopropyltriethoxysilane
(APTES) cathode treated interfacially provides large-area modules with high spatial performance
uniformity (Luck et al. 2013).
Vasilopoulou et al. (2014a) used poly(3-hexylthiophene (P3HT) as the electron donor and
[6,6]-phenyl-C71-butyric acid methyl ester (PC71BM) as the electron acceptor to prepare inverted
OPV cells. ZnO films were used as electron collection layers to improve its performance, while an
under-stoichiometric molybdenum oxide MoOx was used as the hole collection layer. The ZnO layer
by atomic layer deposition (ALD-ZnO) with 20  nm thickness exhibited significant enhancement
in the performance of OPV cells as compared to ZnO deposited by the sol-gel method (sg-ZnO)
in inverted organic solar cells. The ZnO layer deposited by atomic layer deposition show reduced
defect/trap concentration and a decrease in the electron extraction barrier at the ALD-ZnO/organic
active layer interface.
Kim et al. (2014d) introduced ytterbium (Yb) as the electron transport layer in inverted OPVs
where Yb was evaporated directly onto indium tin oxide (ITO). This electron transport layer (ETL)
increases the power conversion efficiency and air stability in organic solar cells by promoting the
formation of ohmic contact between the active layer and the cathode metal. The inverted OPVs
made up of ITO/Yb/P3HT:PCBM/MoO3/Ag give PCE up to 4.3% and fill factor of 71% in one sun
irradiation. The Yb in inverted OPVs is vastly superior to other ETLs. Over 80% of its original PCE
was retained, even after 30 days. Giansante et al. (2014) developed a solution-phase ligand exchange
method, in which they replaced oleate ligands on PbS QDs with arenethiolate anions ligands in
hybrid organic and inorganic quantum dot photovoltaic cells. The objective of using arenethiolate
anions in place of oleate ligands was to decrease the distance between particles in PbS QD solids,
which leads to enhancement of charge transport. This showed an efficiency of 1.85%, which is out-
standing among single PbS QD layers.
Organic photovoltaics have wide applications because they offer many advantages, including
that they are lightweight, thin, semitransparent, mechanically flexible, low cost, and not a disposal
problem. There is an added advantage in that there are no restrictions on the shape, size, and color
(or semitransparent) of OPV devices; therefore, they can be applied to surfaces of nearly any shape.
The use of low-bandgap polymers, small molecules, or a combination of both may enhance the effi-
ciency of OPV cells. An interesting example is that an OPV-functionalized bag is useful in charging
all the portable electronic devices carried in it.

REFERENCES
Adams, W. G., and R. E. Day. 1877. The action of light on selenium. Phil. Trans. Royal Soc. London. 167:
313–349.
Antohe, S., and L. Tugulea. 1991. Electrical and photovoltaic properties of a two-layer organic photovoltaic
cell. Phys. Status Solidi (A). 128: 253–260.
Appleyard, S. F. J., S. R. Day, R. D. Pickford, and M. R. Willis. 2000. Organic electroluminescent devices:
Enhanced carrier injection using SAM derivatized ITO electrodes. J. Mater. Chem. 10: 169–173.
Aziz, F., Z. Ahmad, S. M. Abdullah, K. Sulaiman, and M. H. Sayyad. 2015. Photovoltaic effect in single-
junction organic solar cell fabricated using vanadyl phthalocyanine soluble derivative. Pigment Resin
Technol. 44: 26–32.
Bakulin, A. A., A. Rao, V. G. Pavelyev, P. H. M. van Loosdrecht, M. S. Pshenichnikov, D. Niedzialek, et al.
2012. The role of driving energy and delocalized states for charge separation in organic semiconductors.
Science. 335: 1340–1344.
Becquerel, A. E. 1839. Recherchessur les effets de la radiation chimique de la lumieresolaire au moyen des
courants electriques. Compt. Rend. Acad. Sci. 9: 145–149.
Organic Photovoltaic Cells 79

Beiley, Z. M., M. G. Christoforo, P. Gratia, A. R. Bowring, P. Eberspacher, G. Y. Margulis, et al. 2013. Semi-
transparent polymer solar cells with excellent sub-bandgap transmission for third generation photovolta-
ics. Adv. Mater. 25: 7020–7026.
Beliatis, M. J., K. K. Gandhi, L. J. Rozanski, R. Rhodes, L. McCafferty, M. R. Alenezi, et al. 2014. Hybrid gra-
phene-metal oxide solution processed electron transport layers for large area high-performance organic
photovoltaics. Adv. Mater. 26: 2078–2083.
Bergmann, L. 1931. Uber eine neue Selen-sperrschicht photozelle. Physik. Z. 32: 286–289.
Bernardo, B., D. Cheyns, B. Verreet, R. D. Schaller, B. P. Rand, and N. C. Giebink. 2014. Delocalization and
dielectric screening of charge transfer states in organic photovoltaic cells. Nat. Commun. 5: 3245.
Brabec, C. J., N. S. Sariciftci, and J. C. Hummelen. 2001a. Plastic solar cells. Adv. Funct. Mater. 11: 15–26.
Brabec, C. J., C. Winder, M. C. Scharber, N. S. Sariciftci, J. C. Hummelen, M. Svensson, et al. 2001b. Influence
of disorder on the photo-induced excitations in phenyl substituted polythiophenes. J. Chem. Phys. 115:
7235–7244.
Cacialli, F., R. H. Friend, C.-M. Bouche, P. Le Barny, H. Facoetti, F. Soyer, et al. 1998. Napthalimide side
chain polymers for organic eight-emitting diodes: Band offset engineering and role of polymer thick-
ness. J. Appl. Phys. 83: 2343–2356.
Cahen, D., G. Hodes, M. Gratzel, J. F. Guillemoles, and I. Riess. 2000. Nature of photovoltaic action in dye-
sensitized solar cells. J. Phys. Chem. B. 104: 2053–2059.
Chamberlain, G. A. 1983. Organic solar cells: A review. Solar Cells. 8: 47–83.
Chan, C.-Y., Y.-C. Wong, H.-L. Wong, M.-Y. Chan, and W.-W. Yam. 2014. A new class of three-dimensional,
p-type, spirobifluorene-modified perylene diimide derivatives for small molecular-based bulk hetero-
junction organic photovoltaic devices. J. Mater. Chem. C. 2: 7656–7665.
Chang, D. W., H.-J. Choi, A. Filer, and J.-B. Baek. 2014. Graphene in photovoltaic applications: Organic
photovoltaic cells (OPVs) and dye-sensitized solar cells (DSSCs). J. Mater. Chem. A. 2: 12136–12149.
Chasteen, S. V., V. Sholinb, S. A. Carterb, and Garry Rumblesc. 2008. Towards optimization of device per-
formance in conjugated polymer photovoltaics: Charge generation, transfer and transport in poly(p-
phenylene-vinylene) polymer heterojunctions. Sol. Energy Mater. Solar Cells. 92: 651–659.
Chauhan, R. N., R. S. Anand, and J. Kumar. 2014. RF-sputtered Al-doped ZnO thin films: Optoelectrical
properties and application in photovoltaic devices. Phys. Status Solidi (a). 211: 2514–2522.
Chen, C.-P., Y.-D. Chen, and S.-C. Chuang. 2011. High-performance and highly durable inverted organic pho-
tovoltaics embedding solution-processable vanadium oxides as an interfacial hole-transporting layer.
Adv. Mater. 23: 3859–3863.
Chen, G.-Y., Y.-H. Cheng, Y.-J. Chou, M.-S. Su, C.-M. Chen, and K.-H. Wei. 2011a. Crystalline conjugated
polymer containing fused 2,5-di(thiophen-2-yl)thieno[2,3-b] thiophene and thieno[3,4-c]pyrrole-4,6-di-
one units for bulk heterojunction solar cells. Chem. Commun. 47: 5064–5066.
Chen, Y. C., C. Y. Hsu, R. Y. Lin, K. C. Ho, and J. T. Lin. 2013. Materials for the active layer of organic pho-
tovoltaics: Ternary solar cell approach. ChemSusChem. 6: 20–35.
Cheyns, D., M. Kim, B. Verreet, and B. P. Rand. 2014. Accurate spectral response measurements of a complemen-
tary absorbing organic tandem cell with fill factor exceeding the subcells. Appl. Phys. Lett. 104: 093302.
Chiang, C. K., C. R. Fincher, Jr., Y. W. Park, A. J. Heeger, H. Shirakawa, E. J. Louis, et al. 1977. Electrical
conductivity in doped polyacetylene. Phys. Rev. Lett. 39: 1098.
Cho. J. M., J. Kim, H. Kim, M. Kim, S.-J. Moon, and W. S. Shin. 2014. ITO/AZO double-layered transparent
conducting oxide films for organic photovoltaic cells. Mol. Cryst. Liq. Cryst. 597: 1–7.
Chou, C.-H., W. L. Kwan, Z. Hong, L.-M. Chen, and Y. Yang. 2011. A metal-oxide interconnection layer for
polymer tandem solar cells with an inverted architecture. Adv. Mater. 23: 1282–1286.
Cimrova, V., M. Remmers, D. Neher, and G. Wegner. 1996. Polarized light emission from LEDs prepared by
the Langmuir-Blodgett technique. Adv. Mater. 8: 146–149.
Coakley, K. M., and M. D. McGehee. 2003. Photovoltaic cells made from conjugated polymers infiltrated into
mesoporous titania. Appl. Phys. Lett. 83: 3380.
Coakley, K. M., and M. D. McGehee. 2004. Conjugated polymer photovoltaic cells. Chem. Mater. 16:
4533–4542.
Deibel, C., T. Strobell, and V. Dyakonov. 2010. Role of the charge transfer state in organic donor–acceptor
solar cells. Adv. Mater. 22: 4097–4111.
Dhanabalan, A. J., K. J. Van Duren, P. A. Van Hal, J. L. J. van Dongen, and R. A. J. Janssen. 2001. Synthesis
and characterization of a low bandgap conjugated polymer for bulk heterojunction photovoltaic cells.
Adv. Funct. Mater. 11: 255–262.
Diaz, F. R., J. C. Bernède, M. A. Del Valle, and V. Jousseaume. 2001. From the organic electroluminescent
diodes to the new organic photovoltaic cells. Curr. Trends Polym. Sci. 6: 135–155.
80 Solar Energy Conversion and Storage

Dittmer, J. J., E. A. Marseglia, and R. H. Friend. 2000. Electron trapping in dye/polymer blend photovoltaic
cells. Adv. Mater. 12: 1270–1274.
Dou, L., J. You, J. Yang, C.-C. Chen, Y. He, S. Murase, et al. 2012. Tandem polymer solar cells featuring a
spectrally matched low-bandgap polymer. Nat. Photonics. 6: 180–185.
Elsawy, W., H. Kang, K. Yu, A. Elbarbary, K. Lee, and J.-S. Lee. 2014. Synthesis and characterization
of isoindigo-based polymers using CH-arylation polycondensation reactions for organic photovoltaics.
J. Polym. Sci. Part A. 52: 2926–2933.
Fanshun, M., and H. Tian. 2005. Solar cells based on cyanine and polymethine dyes. In N. S. Sariciftci and
S. S. Sun (Eds), Organic Photovoltaics: Mechanisms, Materials and Devices. Boca Raton: CRC Press.
Fritts, C. E. 1883. On a new form of selenium photocell. Am. J. Sci. 26: 465–472.
Gautier-Thianche, E., C. Sentein, A. Lorin, C. Denis, P. Raimond, and J. M. Nunzi. 1998. Effect of coumarin
on blue light-emitting diodes based on carbazol polymers. J. Appl. Phys. 83: 4236–4241.
Gelinas, S., A. Rao, A. Kumar, S. L. Smith, A. W. Chin, J. Clark, et al. 2014. Ultrafast long-range charge
separation in organic semiconductor photovoltaic diodes. Science. 343: 512–516.
Ghosh, A. K., D. L. Morel, T. Feng, R. F. Shaw, and C. A. Rowe, Jr. 1974. Photovoltaic and rectification proper-
ties of Al/Mg phthalocyanine/Ag Schottky-barrier cells. J. Appl. Phys. 45: 230–236.
Giansante, C., L. Carbone, C. Giannini, D. Altamura, Z. Ameer, G. Maruccio, et al. 2014. Surface chemistry
of arenethiolate-capped PbS quantum dots and application as colloidally stable photovoltaic ink. Thin
Solid Films. 560: 2–9.
Gilchrist, J. B., T. H. Basey-Fisher, S. C. Chang, F. Scheltens, D. W. Mccomb, and S. Heutz. 2014. Uncovering
buried structure and interfaces in molecular photovoltaics. Adv. Funct. Mater. 24: 6473–6483.
Gilot, J., M. M. Wienk, and R. A. J. Janssen. 2010. Optimizing polymer tandem solar cells. Adv. Mater. 22:
E67–E71.
Grondahl, L. O. 1933. The copper-cuprous-oxide rectifier and photoelectric cell. Rev. Modern Phys. 5: 141–168.
Gwang, H. J., S. J. Jin, S. H. Park, S. Jeon, and S. H. Hong. 2012. Highly dispersed carbon nanotubes in
organic media for polymer: Fullerene photovoltaic devices. Carbon. 50: 40–46.
Hadipour, A., B. De Boer, and P. W. M. Blom. 2008. Organic tandem and multi-junction solar cells. Adv.
Funct. Mater. 18: 169–181.
Hadipour, A., B. De Boer, J. Wildeman, F. B. Kooistra, J. C. Hummelen, M. G. R. Turbiez, et al. 2006.
Solution-processed organic tandem solar cells. Adv. Funct. Mater. 16: 1897–1903.
Hai, J., W. Yu, E. Zhu, L. Bian, J. Zhang, and W. Tang. 2014. Synthesis and photovoltaic characterization of
thiadiazole based low bandgap polymers. Thin Solid Films. 562: 75–83.
Halim, M., J. N. G. Pillow, I. D. W. Samuel, and P. L. Burn. 1999. Conjugated dendrimers for light-emitting
diodes: Effect of generation. Adv. Mater. 11: 371–374.
Halls, M., J. J. N. K. Pichler, R. H. Friend, S. C. Morttigard, and A. B. Holmes. 1996. Excitons diffusion and
dissociation in a poly(p-phenylene vinylene)/C60 heterojunction photovoltaic cell. Appl. Phys. Lett. 68:
3120–3122.
Harald, H., and N. S. Sariciftci. 2004. Organic solar cells: An overview. J. Mater. Res. 19: 1924–1945.
He, Z., C. Zhong, S. Su, M. Xu, H. Wu, and Y. Cao. 2012. Enhanced power-conversion efficiency in polymer
solar cells using an inverted device structure. Nat. Photonics 6: 591–595.
Hedley, G. J., A. J. Ward, A. Alekseev, C. T. Howells, E. R. Martins, L. A. Serrano, et al. 2013. Determining
the optimum morphology in high-performance polymer-fullerene organic photovoltaic cells. Nat.
Commun. 4: 2867. doi:10.1038/ncomm53867.
Hendrickx, E., Y. Zhang, K. B. Ferrio, J. A. Herlocker, J. Anderson, N. R. Armstrong, et al. 1999.
Photoconductive properties of PVK-based photorefractive polymer composites doped with fluorinated
styrene chromophores. J. Mater. Chem. 9: 2251–2258.
Hoegel, H. 1965. On photoelectric effects in polymers and their sensitization by dopants. J. Phys. Chem. 69:
755–766.
Jailaubekov, A. E., A. P. Willard, J. R. Tritsch, W.-L Chan, N. Sai, R. Gearba, et al. 2013. Hot charge-transfer
excitons set the time limit for charge separation at donor/acceptor interfaces in organic photovoltaics.
Nat. Mater. 12: 66–73.
Jenekhe, S. A., and S. Yi. 2000. Efficient photovoltaic cells from semiconducting polymer. Appl. Phys. Lett.
77: 2635–2637.
Jiang, J.-M., P.-A. Yang, H.-C. Chen, and K.-H. Wei. 2011. Synthesis, characterization, and photovoltaic prop-
erties of a low-band gap copolymer based on 2,1,3-benzooxadiazole. Chem. Commun. 47: 8877–8879.
Jiang, X., R. A. Register, K. A. Killeen, M. E. Thompson, F. Pschenitzka, and J. C. Sturm. 2000. Statistical
copolymers with side-chain hole and electron transport groups for single layer electroluminescent
device application. Chem. Mater. 12: 2542–2549.
Organic Photovoltaic Cells 81

Jin, G., H.-T. Wei, T.-Y. Na, H.-Z. Sun, H. Zhang, and B. Yang. 2014. High-efficiency aqueous-processed hybrid
solar cells with an enormous Herschel infrared contribution. ACS Appl. Mater. Interfaces. 6: 8606–8612.
Jo, J. W., J. W. Jung, H.-W. Wang, P. Kim, T. P. Russell, and W. H. Jo. 2014. Fluorination of polythiophene
derivatives for high performance organic photovoltaics. Chem. Mater. 26: 4214– 4220.
Karak, S., P. J. Homnick, A. M. D. Pelle, Y. Bae, V. V. Duzhko, F. Liu, et al. 2014. Crystallinity and mor-
phology effect on a solvent-processed solar cell using a triarylamine-substituted squaraine. ACS Appl.
Mater. Interfaces. 6: 11376–11384.
Kido, J., K. Hongawa, K. Okuyama, and K. Nagai. 1994. White light-emitting organic electroluminescent
devices using the poly(N-vinylcarbazole) emitter layer doped with three fluorescent dyes. Appl. Phys.
Lett. 64: 815–817.
Kim, G. M., I. S. Oh, A. N. Lee, and S. Y. Oh. 2014d. Applications of ytterbium in inverted organic photovol-
taic cells as high-performance and stable electron transport layers. J. Mater. Chem. A. 2: 10131–10136.
Kim, J., K. G. Chitibabu, M. J. Cazeca, W. Kim, J. Kumar, and S. K. Tripathy. 1997. Fabrication of polymer
light-emitting diodes by layer-by-layer complexation technique In optical, and magnetic properties of
organic solid-state materials V. Mater. Res. Soc. Symp. Proc. 488: 527–532.
Kim, J.-H., S. A. Shin, J. B. Park, C. E. Song, W. S. Shin, H. Yang, Y. Li, et al. 2014c. Fluorinated benzosele-
nadiazole-based low-band-gap polymers for high efficiency inverted single and tandem organic photo-
voltaic cells. Macromolecules. 47: 1613–1622.
Kim, J.-H., C. E. Song, N. Shin, H. Kang, S. Wood, I.-N. Kang, et al. 2013. High-crystalline medium-band-gap
polymers consisting of benzodithiophene and benzotriazole derivatives for organic photovoltaic cells.
ACS Appl. Mater. Interfaces. 5: 12820–12831.
Kim, J. S., Z. Fei, S. Wood, D. T. James, M. Sim, K. Cho, et al. 2014b. Germanium- and silicon-substituted
donor-acceptor type copolymers: Effect of the bridging heteroatom on molecular packing and photovol-
taic device performance. Adv. Energy Mater. 4. doi :10.1002/aenm.201400527.
Kim, J. Y., K. Lee, N. E. Coates, D. Moses, T.-Q. Nguyen, M. Dante, et al. 2007. Efficient tandem polymer
solar cells fabricated by all-solution processing. Science. 317: 222–225.
Kim, Y., S. Cook, S. M. Tuladhar, S. A. Choulis, J. Nelson, J. R. Durrant, et al. 2006. A strong regioregularity
effect in self-organizing conjugated polymer films and high-efficiency polythiophene:fullerene solar
cells. Nat. Mater. 5: 197–203.
Kim, Y., C. E. Song, S.-J. Moon, and E. Lim. 2014a. Rhodanine dye-based small molecule acceptors for
organic photovoltaic cells. Chem. Commun. 50: 8235–8238.
Ko, D.-H., J. R. Tumbleston, W. Schenck, R. Lopez, and E. T. Samulski. 2011. Photonic crystal geometry for
organic polymer: Fullerene standard and inverted solar cells. J. Phys. Chem. C. 115: 4247–4254.
Ko, D.-H., J. R. Tumbleston, L. Zhang, S. Williams, J. M. DeSimone, R. Lopez, et al. 2009. Photonic crystal
geometry for organic solar cells. Nano Lett. 9: 2742–2746.
Kroon, R., A. Diaz De Zerio Mendaza, S. Himmelberger, J. Bergqvist, O. Bäcke, G. C. Faria, et al. 2014. A
new tetracyclic lactam building block for thick, broad-bandgap photovoltaics. J. Am. Chem. Soc. 136:
11578–11581.
Le, Q. V., T. P. Nguyen, H. W. Jang, and S. Y. Kim. 2014. The use of UV/ozone-treated MoS2 nanosheets for
extended air stability in organic photovoltaic cells. Phys. Chem. Chem. Phys. 16: 13123–13128.
Lee, C.-H., R. Pandey, B.-Y. Wang, W.-K. Choi, D.-K. Choi, and Y.-J. Oh. 2015. Nano-sized indium-free MTO/
Ag/MTO transparent conducting electrode prepared by RF sputtering at room temperature for organic
photovoltaic cells. Sol. Energy Mater. Sol. Cells. 132: 80–85.
Lee, D. Y., D. V. Shinde, S. J. Yoon, K. N. Cho, W. Lee, N. K. Shrestha, et al. 2014. Cu-based metal–organic
frameworks for photovoltaic application. J. Phys. Chem. C. 118: 16328–16334.
Lee, S.-K., K. Rana, and J.-H. Ahn. 2013. Graphene films for flexible organic and energy storage devices.
J. Phys Chem. Lett. 4: 831–841.
Li, G., V. Shrotriya, J. Huang, Y. Tao, T. Moriarty, K. Emery, et al. 2005. High-efficiency solution processable
polymer photovoltaic cells by self-organization of polymer blends. Nat. Mater. 4: 864–868.
Li, N., D. Baran, K. Forberich, M. Turbiez, T. Ameri, F. C. Krebs, and C. J. Brabec. 2013. An efficient solu-
tion-processed intermediate layer for facilitating fabrication of organic multi-junction solar cells. Adv.
Energy Mater. 3: 1597–1605.
Lim, K.-G., M.-R. Choi, J.-H. Kim, D. H. Kim, G. H. Jung, Y. Park, et al. 2014. Role of ultrathin metal fluoride
layer in organic photovoltaic cells: Mechanism of efficiency and lifetime enhancement. ChemSusChem.
7: 1125–1132.
Luck, K. A., T. A. Shastry, S. Loser, G. Ogien, T. J. Marks, and M. C. Hersam. 2013. Improved uniformity in
high-performance organic photovoltaics enabled by (3-aminopropyl)triethoxysilane cathode function-
alization. Phys. Chem. Chem. Phys. 15: 20966–20972.
82 Solar Energy Conversion and Storage

Marks, R. N., J. J. M. Halls, D. D. C. Bradley, R. H. Friend, and A. B. Holmes. 1994. The photovoltaic response
in poly(p-phenylene vinylene) thin-film devices. J. Phys.: Condens. Matter. 6: 1379–1394.
McGehee. M. D., and M. A. Topinka. 2006. Solar cells: Pictures from the blended zone. Nat. Mater. 5:
675–676.
Miller, S., G. Fanchini, Y.-Y. Lin, C. Li, C.-W. Chen, W.-F. Su, et al. 2008. Investigation of nanoscale morpho-
logical changes in organic photovoltaics during vapour annealing. J. Mater. Chem. 18: 306–312.
Mitul, A. F., L. Mohammad, S. Venkatesan, N. Adhikari, S. Sigdel, Q. Wang, et al. 2015. Low temperature
efficient interconnecting layer for tandem polymer solar cells. Nano Energy. 11: 56–63.
Nelson, J. 2002. Organic photovoltaic films. Curr. Opin. Solid State Mater. Sci. 6: 87–95.
Nielsen, C. B., E. Voroshazi, S. Holliday, K. Cnops, D. Cheyns, and I. McCulloch. 2014. Electron-deficient
truxenone derivatives and their use in organic photovoltaics. J. Mater. Chem. A. 2: 12348–12354.
Nikiforov, M. P., J. Strzalka, Z. Jiang, and S. B. Darling. 2013. Lanthanides: New metallic cathode materials
for organic photovoltaic cells. Phys. Chem. Chem. Phys. 15: 13052–13060.
Nix, F. C., and A. W. Treptwo. 1939. A thallous sulphide photo EMF cell. J. Opt. Soc. Am. 29: 457–462.
Nuesch, F., L. Si-Ahmed, B. Francois, and L. Zuppiroli. 1997. Derivatized electrodes in the construction of
organic light emitting diodes. Adv. Mater. 9: 222–225.
Owczarczyk, Z. R., W. A. Braunecker, S. D. Oosterhout, N. Kopidakis, R. E. Larsen, D. S. Ginley, et al. 2014.
Cyclopenta[c]thiophene-4,6-dione-based copolymers as organic photovoltaic donor materials. Adv.
Energy Mater. 4. doi: 10.1002/ aenm.201301821.
Paci, B., A. Generosi, V. R. Albertini, and R. De Bettignes. 2013. Improved structural/morphological durabil-
ity for organic solar cells based on poly(3-hexylthiophene) fibers photoactive layers. Chem. Phys. Lett.
587: 50–55.
Page, Z. A., Y. Liu, V. V. Duzhko, T. P. Russell, and T. Emrick. 2014. Fulleropyrrolidine interlayers: Tailoring
electrodes to raise organic solar cell efficiency. Science. 346: 441–444.
Park, Y. S., T. S. Kale, C.-Y. Nam, D. Choi, and R. B. Grubbs. 2014. Effects of heteroatom substitution in
conjugated heterocyclic compounds on photovoltaic performance: From sulfur to tellurium. Chem.
Commun. 50: 7964–7967.
Petritsch, K., J. J. Dittmer, E. A. Marseglia, R. H. Friend, A. Lux, G. G. Rozenberg, et al. 2000. Dye based
donor/acceptor solar cells, Sol. Energy Mater. Solar Cells. 61: 63–72.
Petritsch, K., R. H. Friend, A. Lux, G. Rozenberg, S. C. Moratti, and A. B. Holmes. 1999. Liquid crystalline
phthalocyanines in organic solar cells. Synth. Met. 102: 1776–1777.
Peumans, P., A. Yakimov, and S. R. Forrest. 2003. Small molecular weight thin-film photodetectors and solar
cells. J. Appl. Phys. 93: 3693–3723.
Pfannmoller, M., W. Kowalsky, and R. R. Schroder. 2013. Visualizing physical, electronic, and optical proper-
ties of organic photovoltaic cells. Energy Environ. Sci. 6: 2871–2891.
Piliegoa, C., and M. A. Loi. 2012. Application charge transfer state in highly efficient polymer-fullerene bulk
heterojunction solar cells. J. Mater. Chem. 22: 4141–4150.
Piliego, C., T. W. Holcombe, J. D. Douglas, C. H. Woo, P. M. Beaujuge, and J. M. J. Fréchet. 2010. Synthetic
control of structural order in N-alkylthieno[3,4-c]pyrrole-4,6-dione-based polymers for efficient solar
cells. J. Am. Chem. Soc. 132: 7595–7597.
Pochettino, A. 1906. Sul comportamento foto-elettrico dell’antracene. Acad. Lincei Rend. 15: 355–368.
Riede, M., C. Uhrich, J. Widmer, R. Timmreck, D. Wynands, G. Schwartz, et al. 2011. Efficient organic tan-
dem solar cells based on small molecules. Adv. Funct. Mater. 21: 3019–3028.
Roman, L. S., A. C. Arias, M. Theander, M. R. Andersson, and O. Inganas. 2003. Photovoltaic devices based
on photo induced charge transfer in polythiophene:CN-PPV blends. Braz. J. Phys. 33: 376–381.
Roman, L. S., L. A. A. Patterson, and O. J. Inganas. 1999. Modeling photocurrent action spectra of photovol-
taic devices based on organic thin films. J. Appl. Phys. 86: 487–496.
Salafsky, J. S. 1999. Exciton dissociation, charge transport, and recombination in ultrathin, conjugated poly-
mer–TiO2 nanocrystal intermixed composites. Phys. Rev. B. 59: 10885–10894.
Saliba, M., K. W. Tan, H. Sai, D. T. Moore, T. Scott, W. Zhang, et al. 2014. Influence of thermal process-
ing protocol upon the crystallization and photovoltaic performance of organic-inorganic lead trihalide
perovskites. J. Phys. Chem. C. 118: 17171–17177.
Sariciftci, N. S. 2004. Plastic photovoltaic devices. Mater. Today. 7: 36–40.
Sariciftci, N. S., D. Braun, C. Zhang, V. I. Srdanov, A. S. Heeger, G. Stucky, et al. 1993. Semiconducting
polymer-buckminsterfullerene heterojunctions: Diodes, photodiodes, and photovoltaic cells. Appl.
Phys. Lett. 62: 585–587.
Sariciftci. N. S., L. Smilowitz, A. J. Heeger, and F. Wudl. 1992. Photoinduced electron transfer from a con-
ducting polymer to buckminsterfullerene. Science. 258: 1474–1476.
Organic Photovoltaic Cells 83

Scharber, M. C., D. Mühlbacher, M. Koppe, P. Denk, C. Waldauf, A. J. Heeger, and C. J. Brabec. 2006. Design
rules for donors in bulk-heterojunction solar cells—Towards 10% energy-conversion efficiency. Adv.
Mater. 18: 789–794.
Schmidt-Mende, L., A. Fechtenkötter, K. Müllen, E. Moons, R. H. Friend, and J. D. MacKenzie. 2001. Self-
organized discotic liquid crystals for high-efficiency organic photovoltaics. Science. 293: 1119–1122.
Schon, J. H., K. Ch, and B. Batlogg. 2000. Efficient photovoltaic energy conversion in pentacene-based het-
erojunction. Appl. Phys. Lett. 77: 2473–2475.
Seguy, I., P. Destruel, and H. Bock. 2000. An all-columnar bilayer light-emitting diode. Synth. Met. 15:
111–112.
Seo, H. O., S.-Y. Park, W. H. Shim, K.-D. Kim, K. H. Lee, M. Y. Jo, et al. 2011. Ultrathin TiO2 films on ZnO
electron-collecting layers of inverted organic solar cell. J. Phys. Chem. C. 115: 21517–21520.
Shaheen, S. E., C. J. Brabec, N. S. Sacriciftci, F. Padinger, T. Fromherz, and J. C. Hummelen. 2001a. 2.5%
Efficient organic plastic solar cell. Appl. Phys. Lett. 78: 841–843.
Shaheen, S. E., D. Vangeneugden, R. Kiebooms, D. Vanderzande, T. Fromherz, F. Padinger, et al. 2001b. Low
band-gap polymeric photovoltaic devices. Synth. Met. 121: 1583–1584.
Shirakawa, H., E. J. Louis, A. G. MacDiarmid, C. K. Chiang, and A. J. Heeger. 1977. Synthesis of electri-
cally conducting organic polymers: Halogen derivatives of polyacetylene, (CH)x. J. Chem. Soc. Chem.
Commun. 578–580.
Shivanna, R., S. Shoaee, S. Dimitrov, S. K. Kandappa, S. Rajaram, J. R. Durrant, et al. 2014. Charge genera-
tion and transport in efficient organic bulk heterojunction solar cells with a perylene acceptor. Energy
Environ. Sci. 7: 435–441.
Sista, S., M.-H. Park, Z. Hong, Y. Wu, J. Hou, W. L. Kwan, G. Li, and Y. Yang. 2010. Highly efficient tandem
polymer photovoltaic cells. Adv. Mater. 22: 380–383.
Skupov, K., and A. Adronov. 2014. Effect of carbon nanotube incorporation into polythiophene-fullerene-
based organic solar cells. Can. J. Chem. 92: 68–75.
Smith, W. 1873. Effect of light on selenium during the passage of an electric current. Nature. 7: 303.
Stanculescu, F., O. Rasoga, A. M. Catargiu, L. Vacareanu, M. Socol, C. Breazu, et al. 2014. MAPLE prepared
heterostructures with arylene based polymer active layer for photovoltaic applications. Appl. Surface
Sci. 336: 240–248.
Sun, Y., J. H. Seo, C. J. Takacs, J. Seifter, and A. J. Heeger. 2011. Inverted polymer solar cells integrated with a low-
temperature-annealed sol-gel-derived ZnO film as an electron transport layer. Adv. Mater. 23: 1679–1683.
Takao, Y., T. Masuoka, K. Yamamoto, T. Mizutani, F. Matsumoto, K. Moriwaki, et al. 2014. Synthesis and
properties of novel fluorinated subnaphthalocyanine for organic photovoltaic cells. Tetrahedron Lett.
55: 4564–4567.
Tamura, Y., H. Saeki, J. Hashizume, Y. Okazaki, D. Kuzuhara, M. Suzuki, et al. 2014. Direct comparison of a
covalently-linked dyad and a 1:1 mixture of tetrabenzoporphyrin and fullerene as organic photovoltaic
materials. Chem. Commun. 50: 10379–10381.
Tan, W.-Y., R. Wang, M. Li, G. Liu, P. Chen, X.-C. Li, et al. 2014. Lending triarylphosphine oxide to phenan-
throline: A facile approach to high-performance organic small-molecule cathode interfacial material for
organic photovoltaics utilizing air-stable cathodes. Adv. Funct. Mater. 24: 6540–6547.
Tang, C. W. 1986. Two-layer organic photovoltaic cell. Appl. Phys. Lett. 48: 183–185.
Theander, M., A. Yartsev, D. Zigmantas, V. Sundstrom, W. Mammo, M. R. Andersson, et al. 2000.
Photoluminescence quenching at a polythiophene/C60 heterojunction. Phys. Rev. B. 61: 12957–12963.
Thompson, B. C., and J. M. J. Fréchet. 2008. Polymer-fullerene composite solar cells. Angew. Chem. Int. Ed.
47: 58–77.
Tokuhisa, H., M. Era, and T. Tsutsui. 1998. Polarized electroluminescence from smectic mesophase. Appl.
Phys. Lett. 72: 2639–2641.
Too, C. O., G. G. Wallace, A. K. Burrell, G. E. Collis, D. L. Officer, E. W. Boge, et al. 2001. Photovoltaic
devices based on polythiophenes and substituted polythiophenes. Synth. Met. 123: 53–60.
Tsai, C.-C., R. R. Grote, J. H. Beck. I. Kymissis, R. M. Osgood, Jr., and D. Englund. 2014. General method for
simultaneous optimization of light trapping and carrier collection in an ultra-thin film organic photovol-
taic cell. J. Appl. Phys. 116: 023110. doi:10.1063/1.4890275.
Tung, V. C., J. Kim, L. J. Cote, and J. Huang. 2011. Sticky interconnect for solution-processed tandem solar
cells. J. Am. Chem. Soc. 133: 9262–9265.
Tvingstedt, K., V. Andersson, F. Zhang, and O. Inganas. 2007. Folded reflective tandem polymer solar cell
doubles efficiency. Appl. Phys. Lett. 91: 123514. doi:10.1063/1.2789393.
Umeyama, T., and H. Imahori. 2014. Design and control of organic semiconductors and their nanostructures
for polymer-fullerene-based photovoltaic devices. J. Mater. Chem. A. 2: 11545–11560.
84 Solar Energy Conversion and Storage

Vasilopoulou, M., D. G. Georgiadou, A. Soultati, N. Boukos, S. Gardelis, L. C. Palilis, et al. 2014b. Atomic-
layer-deposited aluminum and zirconium oxides for surface passivation of TiO2 in high-efficiency
organic photovoltaics. Adv. Energy Mater. 4. doi: 10.1002/ aenm.201400214.
Vasilopoulou, M., N. Konofaos, D. Davazoglou, P. Argitis, N. A. Stathopoulos, S. P. Savaidis, et al. 2014a.
Organic photovoltaic performance improvement using atomic layer deposited ZnO electron-collecting
layers. Solid-State Electron. 101: 50–56.
Walter, M. G., A. B. Rudine, and C. C. Wamser. 2010. Porphyrins and phthalocyanines in solar photovoltaic
cells. J. Porphyrins Phthalocyanines. 14: 759–792.
Wang, S., S. Yang, C. Yang, Z. Li, J. Wang, and W. Ge. 2000. Poly(N-vinylcarbazole) (PVK) photoconductiv-
ity enhancement induced by doping with CdS nanocrystals through chemical hybridization. J. Phys.
Chem. B. 104: 11853–11858.
Wang, T., K. C. Weerasinghe, D. Lui, W. Li, X. Yan, X. Zhou, and L. Wang. 2014b. Ambipolar organic semi-
conductors with cascades of energy levels for generating long-lived charge separated states: A donor-
acceptor1-acceptor2 architectural triarylamine dye. J. Mater. Chem. C. 2: 5466–5470.
Wang, X., J. Huang, Z. Niu, X. Zhang, Y. Sun, and C. Zhan. 2014a. Dimeric naphthalene diimide based
small molecule acceptors: Synthesis, characterization, and photovoltaic properties. Tetrahedron. 70:
4726–4731.
Wang, X., G. I. Koleilat, J. Tang, H. Liu, I. J. Kramer, R. Debnath, et al. 2011. Tandem colloidal quantum dot
solar cells employing a graded recombination layer. Nat. Photon. 5: 480–484.
Wang, Z., T. Miyadera, A. Saeki, Y. Zhou, S. Seki, Y. Shibata, et al. 2014c. Structural influences on charge
carrier dynamics for small-molecule organic photovoltaics. J. Appl. Phys. 116: 013105.
Weinberger, B. R., M. Akhtar, and S. C. Gau. 1982. Polyacetylene photovoltaic devices. Synth. Met. 4: 187–197.
Weiss, N. O., H. Zhou, L. Liao, Y. Liu, S. Jiang, Y. Huang, and X. Duan. 2012. Graphene: An emerging elec-
tronic material. Adv. Mater. 24: 5782–5825.
Wu, A., T. Fujuwara, M. Jikei, M.-A. Kakimoto, Y. Imai, T. Kubota, et al. 1996. Electrical properties and elec-
troluminescence of poly(p-phenylene vinylene) Langmuir-Blodgett film. Thin Solid Films. 284–285:
901–903.
Xia, F., H. Yan, and P. Avouris. 2013. The interaction of light and graphene: Basics, devices, and applications.
Proc. IEEE. 101: 1717–1731.
Yang, J., R. Zhu, Z. Hong, A. Kumar, Y. Li, and Y. Yang. 2011. A robust inter-connecting layer for achieving
high performance tandem polymer solar cells. Adv. Mater. 23: 3465–3470.
Yoshino, K., K. Tada, A. Fujii, E. M. Conwell, and A. A. Zakhidov. 1997. Novel photovoltaic devices based on
donor-acceptor molecular and conducting polymer systems. IEEE Trans. Electron Dev. 44: 1315–1324.
Yu, G., and A. J. Heeger. 1995. Charge separation and photovoltaic conversion in polymer composites with
internal donor/acceptor heterojunctions. J. Appl. Phys. 78: 4510.
Yu, G., J. Gao, J. C. Hummelen, F. Wudl, and A. J. Heeger. 1995. Polymer photovoltaic cells: Enhanced effi-
ciencies via a network of internal donor-acceptor heterojunctions. Science. 270: 1789–1791.
Yu, Y.-Y., and S.-H. Chan. 2013. Effects of metal oxide as an anode interlayer for organic photovoltaics. Thin
Solid Films. 546: 231–235.
Zhang, Q., B. Kan, F, Liu, G. Long, X. Wan, X. Chen, et al. 2015. Small-molecule solar cells with efficiency
over 9%. Nat. Photon. 9: 35–41.
Zhao, D. W., L. Ke, Y. Li, S. T. Tan, A. K. K. Kyaw, H. V. Demir, et al. 2011. Optimization of inverted tandem
organic solar cells. Sol. Energy Mater. Solar Cells. 95: 921–926.
Zhao, D. W., X. W. Sun, C. Y. Jiang, A. K. K. Kyaw, G. Q. Lo, and D. L. Kwong. 2008. Efficient tandem
organic solar cells with an Al/ MoO3 intermediate layer. Appl. Phys. Lett. 93: 083305.
Zhuang, T. X.-F. Wang, T. Sano, Z. Hong, Y. Yang, and J. Kido. 2013. Fullerene derivatives as electron donor
for organic photovoltaic cells. Appl. Phys. Lett. 103: 203301–203304.
Zimmerman, J. D., B. E. Lassiter, X. Xiao, K. Sun, A. Dolocan, R. Gearba, et al. 2013a. Control of interface
order by inverse quasi-epitaxial growth of squaraine/fullerene thin film photovoltaics. ACS Nano. 7:
9268–9275.
Zimmerman, J. D., B. Song, O. Griffith, and S. R. Forrest. 2013b. Exciton-blocking phosphonic acid-treated
anode buffer layers for organic photovoltaics. Appl. Phys. Lett. 103: 243905.
6 Dye-Sensitized Solar Cells
Rakshit Ameta, Surbhi Benjamin, Shweta
Sharma, and Monika Trivedi

CONTENTS
6.1 Introduction............................................................................................................................. 85
6.2 Principle................................................................................................................................... 87
6.3 Characterization...................................................................................................................... 88
6.4 Incident Photon-to-Current Conversion Efficiency................................................................. 89
6.5 Components of DSSC.............................................................................................................. 89
6.6 Sensitizer.................................................................................................................................. 89
6.6.1 Metal Complexes......................................................................................................... 89
6.6.1.1 Ru Metal Complex Dyes...............................................................................90
6.6.1.2 Other Metal Complexes................................................................................90
6.6.2 Dyes............................................................................................................................. 91
6.6.2.1 Metal Complex Free Dyes............................................................................ 91
6.6.3 Natural Pigments.........................................................................................................92
6.7 Electrolytes.............................................................................................................................. 93
6.7.1 Liquid Electrolytes...................................................................................................... 93
6.7.2 Polymer Gel Electrolytes.............................................................................................94
6.7.3 Other Electrolytes........................................................................................................94
6.8 Electrodes................................................................................................................................ 95
6.8.1 Nanocrystalline Semiconductor Working Electrode................................................... 95
6.8.1.1 Naïve Semiconductors.................................................................................. 95
6.8.1.2 Modified Semiconductors.............................................................................97
6.8.1.3 Doped Semiconductors................................................................................. 98
6.8.1.4 Other Nanomaterials.....................................................................................99
6.9 Counterelectrode.................................................................................................................... 100
6.10 Other Types of Dye-Sensitized Solar Cell............................................................................. 101
6.10.1 Solid-State DSSC....................................................................................................... 101
6.10.2 Quasi-Solid-State DSSC............................................................................................ 103
6.10.3 Quantum Dot DSSC.................................................................................................. 104
6.11 Applications of DSSC............................................................................................................ 105
References....................................................................................................................................... 105

6.1 INTRODUCTION
Solar insolation is the most abundant natural energy source for Earth. Solar radiation can be trans-
formed into various energy forms such as heat and chemical energy (e.g., via photosynthesis).
Natural energy sources will be depleted in the near future due to their rapid and uncontrolled
consumption. In this regard, solar energy may play a very important role as an alternative energy
source. The Sun provides about 120,000 TW of energy toward Earth’s surface. This amount is
about 6000 times the present rate of the world’s energy consumption. It is a challenge to use solar
energy as environmentally friendly chemical fuels and photoconversion devices are necessary to

85
86 Solar Energy Conversion and Storage

convert it to electricity. The main advantage of the new generation of solar cells is their low cost
(less than $1/peak watt) and potential (Gratzel 2003).
A solar cell is a solid-state electrical device, which directly converts solar energy into elec-
tricity. Solar energy is transmitted in the form of small packets of photons or quantums of light,
and the generated electrical energy is stored in electromagnetic fields. The solar age began in
the 1950s at Bell Laboratories with the development of silicon technology. They described the
first high-power silicon photovoltaic cell with 6% efficiency. The efficiency of this cell was much
better than the previous selenium solar cell. Due to the energy crisis in 1970, many researchers
developed technologies for producing energy from renewable energy sources and photovoltaic
devices. In 1997, 33% of electrical energy was generated from photovoltaic solar devices, which is
an outstanding achievement.
The first-generation solar cells were single-junction devices. The main problems of this type of
solar cells include that they are labor intensive, require a cumbersome method to produce energy, and
have high manufacturing costs, but their electrical output is relatively high. The second-generation
solar cell was developed to solve this problem. In a thin-film solar cell, cadmium telluride (CdTe),
copper indium gallium selenide (ClGS), copper indium diselenide (CIS), and polycrystalline and
amorphous silicon were applied. Their costs were much lower than those of the previous solar cells.
Then the third-generation solar cells were designed to enhance the electrical performance of thin-
film solar cells, such as polymer solar cells, dye-sensitized solar cells (DSSCs), and nanocrystalline
solar cells.
This new generation of solar cell was based on nanotechnology to lower costs. The mechanism
of conversion of light energy into electrical energy is based on photogeneration of charge carriers
(electrons and holes) on semiconductor material (light-absorbing material) and charge separation
to produce electrical energy. These solar cells are a promising renewable energy source, which are
environmentally friendly based on specific inorganic and organic compounds.
The history of photovoltaics started with the work of French physicist Alexandre-Edmond
Becquerel (1839). He observed that when two platinum electrodes were immersed in an illumi-
nated solution, electric current was generated. The electrodes were coated with light-sensitive metal
halide salt (e.g., AgCl or AgBr). It was observed that as the light intensity was increased, electricity
also increased. The photovoltaic effect in selenium was observed by Smith (1873). Adams and Day
(1876) investigated the electrical behavior of selenium and, especially, its sensitivity to light. They
discovered the photovoltaic effect of the platinum and selenium junction. Moser (1887) discovered
the photosensitization effect on dyes. He used silver halide electrode painted with red dye. Gerischer
et al. (1968) explained the mechanism of dye adsorption on a semiconductor to produce electricity
under certain conditions.
Dye-sensitized electrochemical photocells have been used to convert light energy into electrical
energy. Due to poor light harvesting and the instability of dyes, the efficiency of this device was
very low (i.e., near 1% or less). Matsumura et al. (1980) reported the dye sensitization on ZnO, CdS,
and TiO2 in an electrochemical photocell. They used cationic, anionic, and zwitterionic dyes. A
dye-sensitized photocell with 2.5% cell efficiency was obtained using aluminum-doped porous ZnO
electrode with rose Bengal (anionic dye).
O’Regan and Grätzel (1991) described the heterojunction three-dimensional fabrication of a dye-
sensitized solar cell. The basic components of this device were a semiconductor film containing
nanometer-sized TiO2 particles, with newly developed charge-transfer dyes. They reported more
than 7% efficiency in the presence of direct sunlight. DSSCs are different from conventional semi-
conductor devices where an n-type semiconductor material such as TiO2 was used. It generates
current while dye molecules absorbed photon from light and injected electrons into the conduction
band. DSSCs are photoelectrochemical solar cells. Their mechanism is based on photoinduced
charge separation at a dye-sensitized interface between a nanocrystalline, mesoporous metal oxide
electrode and a redox electrolyte (O’Regan and Durrant 2009).
Dye-Sensitized Solar Cells 87

The DSSCs are also known as “Gratzel cells.” In the last decade, the DSSC has become a
more active and interesting research topic because of its low cost, easy fabrication, low energy
payback time, flexibility, and multicolor option. These cells are technically and economically
as well as environmentally reliable alternatives to existing p-n junction photovoltaic devices.
However, some improvements are still needed in their properties and durability for their future
development.

6.2 PRINCIPLE
A DSSC has some basic steps for generating electrical energy at the cost of light energy. These steps
are

• Nanocrystalline TiO2 is deposited on the conducting electrode (photoelectrode) to provide


a large surface area on which to absorb dye molecules (D) (sensitizer):

D + Photon → D* →Excitation process (6.1)

D* + TiO2 → e− (TiO2 )+D+ → Injection process (6.2)


e− (TiO2 )+ C.E. → TiO2 + e− (C.E.)+ Electrical energy → Energy generation (6.3)


3 1
D+ + I− → D + I3− → Regeneration of dye (6.4)
2 2

1 3
I3− + e− (C.E.) → I− + C.E. → e− Recapture reaction (6.5)
2 2

• When the photons are absorbed, dye molecules are excited from the highest occupied
molecular orbital (HOMO)/valence band (VB) to the lowest unoccupied molecular orbital
(LUMO)/conduction band (CB) states (Equation 6.1).
• Once an electron is injected into the conduction band of the wide-bandgap semiconduc-
tor, like a nanostructured TiO2 film, the excited dye molecules (D*) were oxidized to D+
(Equation 6.2).
• The injected electron is transported between the TiO2 nanoparticles and then extracted to
a load, where the work done is delivered as electrical energy (Equation 6.3).
• Electrolytes containing an I−/I3− redox couple system are used as an electron mediator
between the TiO2 photoelectrode and the carbon-coated counterelectrode. Therefore, the
oxidized dye molecules are regenerated by receiving electrons from the I− ions redox medi-
ator, which is oxidized to I3− (triiodide ions) (Equation 6.4).
• The I3− ion accepts the internally donated electron from the external load and is reduced
back to I− ion (Equation 6.5).

The movement of electrons in the conduction band of the semiconductor is accompanied by the
diffusion of charge-compensating cations in the electrolyte layer close to the nanoparticle surface.
Therefore, generation of electric power in the DSSC causes no permanent chemical change or trans-
formation (Gratzel 2005).
88 Solar Energy Conversion and Storage

6.3 CHARACTERIZATION
The photovoltaic cell is a device that generates electrical power from conversion of incident
light into electrical energy. This process includes producing voltage in the presence of an exter-
nal load and supplying it through any load at the same time. This is presented by the cur-
rent–voltage (i–V) characteristic curve of the cell at different illuminations and temperatures
(Figure 6.1).
When the cell is short-circuited under illumination, the maximum current isc (the short-circuit
current) is generated. Under open-circuit conditions, no current can flow, and the voltage obtained
is at its maximum, which is called the open-circuit voltage (VOC). The point in the i–V curve yield-
ing the maximum product of current and voltage (i.e., power) is called the power point. Here, the
electrical power is maximum (Pmax = imax × Vmax). Another important characteristic of the solar cell
performance is its fill factor (FF), which is defined as

imax ×Vmax
FF = (6.6)
isc ×Voc

where:
imax is the current per unit area at the maximum output power point
Vmax is the voltage per unit area at the maximum output power point
isc is the short-circuit photocurrent density
Voc is the open-circuit voltage
Pin is the light power (intensity of light) per unit area
FF is the fill factor of the cell

The solar energy to electricity conversion efficiency (η) of DSSC is calculated by the following
equation:

isc ×Voc × FF (6.7)


η(%) = ×100
Pin

isc
ipp Power point
Current (i)

Vpp Voc
Voltage (V )

FIGURE 6.1  Current–voltage (i–V) characteristics of the cell.


Dye-Sensitized Solar Cells 89

6.4  INCIDENT PHOTON-TO-CURRENT CONVERSION EFFICIENCY


Incident photon-to-current conversion efficiency (IPCE) is defined as the ratio of the number of
electrons in the external circuit produced by an incident photon at a given wavelength:

100 ×1240 × isc


IPCE (%) = (6.8)
(win × λ)

where:
isc is the short-circuit photocurrent (mA cm−2)
win is the incident light intensity (W cm−2)
λ is wavelength (nm)

6.5  COMPONENTS OF DSSC


The DSSC consists of two electrodes: photoactive working electrode and counterelectrode
(Figure 6.2). These electrodes are contacted by a liquid redox electrolyte. They are explained as
follows:

• Working electrode: Porous nanostructures semiconductor (TiO2) attached to a conducting


fluorine-doped tin oxide (FTO) glass
• Counterelectrode: Platinized conducting substrate
• Light-absorbing layer: Adsorbed dye as a sensitizer
• Redox system: A liquid electrolyte containing the redox couple system (iodide/triiodide)

6.6 SENSITIZER
6.6.1  Metal Complexes
Transition-metal complexes such as derivatives of polypyridine complexes and metalloporphy-
rins were the first choice as photosensitizers due to their low-lying metal-to-ligand charge-transfer
(MLCT) and ligand-centered excited states for maximum electron transfer. Therefore, these can
also be used as photochemical systems for direct conversion and storage of solar energy. The metal

TiO2 electrode Dye Electrolyte Pt electrode


D*/D+
e– Injection
C.B. ∆V
FTO glass

FTO glass

hn – –
I /I3
e–

D+/D
V.B.
E Anode Cathode
e–
e–

e e –
e– –
e

External load

FIGURE 6.2  Dye-sensitized solar cell.


90 Solar Energy Conversion and Storage

complexes attached to the surfaces of some mesoporous membrane thin films with large surface
area were prepared by nanosized colloidal semiconductor dispersions that improve the energy con-
version efficiency of photovoltaic devices such as dye-sensitized solar cells (Kalyanasundaram and
Gratzel 1998).

6.6.1.1  Ru Metal Complex Dyes


Anderson et al. (1979) reported that the attachment of the [Ru(Bipy)3]2+ derivative on the single-crystal
n-type TiO2 electrode produced anodic photocurrents in the presence of visible light. The efficient
sensitizer based on the N-bound isothiocyanato complex K[Ru(bmipy)(dcbpy)(NCS)] on a nano-
crystalline TiO2 surface was suggested by Kohle et al. (1996). A black trithiocyanatoruthenium(II)
terpyridyl complex was found to be quite efficient for the sensitization of the nanocrystalline TiO2
solar cell (Nazeeruddin et al. 1997), whereas Imahori et al. (2009) reported that the ruthenium(II)
bipyridyl complexes were the most efficient complexes in the TiO2-sensitized solar cell. The large
π-aromatic molecules such as porphyrins, phthalocyanines, and perylenes also play an important
role as good sensitizers to give high conversion efficiency in DSSC. Porphyrin and phthalocyanine
sensitized DSSCs have shown poor light-harvesting tendencies compared to ruthenium complexes;
therefore, some types of push–pull porphyrin sensitizers were discovered to further improve power
conversion efficiency to the order of 6%–7%.
The other derivatives of the Ru(II) complex were also used as the sensitizers in DSSCs. Dare-
Edwards et al. (1980) used Ru(bipy)2(bpca)–sensitized n-TiO2, n-SrTiO3, and n-SnO2 electrodes in
solar cells. The N719-sensitized multilayered mono- and double-layered film DSSCs were described
by Wang et al. (2004a). The solar-to-electric energy conversion efficiency has been increased from
7.6% to 9.8% while changing monolayer to multilayer films, and 10.2% efficiency was achieved by
using an antireflection multilayered film in the presence of AM 1.5 solar light (100 mW cm–2).
A new ion-coordinating ruthenium polypridyl sensitizer dye [K68, NaRu(4-carboxylic acid-4′-
carboxylate)(4,4′-bis[(triethyleneglycolmethylether) heptylether]-2,2′-bipyridine)(NCS)2] has been
synthesized and used by Kuang et al. (2008). Binary ionic liquid electrolytes 1-propyl-3-methyl-
imidazolium iodide (PMII) and 1-ethyl-3 methyl-imidazoliumtetracyanoborate (EMIB)(CN)4) have
been used. They achieved 7.7% conversion efficiency for DSSC. The other amphiphilic hetero-
leptic ruthenium(II) sensitizer series, such as [Ru(H2dcbpy)(dhbpy)(NCS)2], [Ru(H2dcbpy)(bccbpy)
(NCS)2], [Ru(H2dcbpy) (mpubpy)(NCS)2], and [Ru(H2dcbpy)(bhcbpy)(NCS)2], have been synthe-
sized for DSSC where the latter sensitizer gave 7.4% efficient cells from all other dyes (Lagref et al.
2008). The DSSC has been sensitized with black dye [(C4H9)4N]3[Ru(Htcterpy)(NCS)3] (tcterpy
= 4,4′,4″-tricarboxy-2,2′,2″-terpyridine), and maximum power conversion efficiency (10.5%) was
achieved from the pretreatment of HCl on the TiO2 surface (Wang et al. 2005a). Ruthenium(II)
complexes containing [Ru(dcbpyH2)2(NCS)2] (N3) and [Ru(dcbpyH2) (tdbpy)(NCS)2] (N621)
dye-absorbed TiO2 films in their monoprotonated and diprotonated states have approached
high power conversion efficiencies of 11.18% and 9.57%, respectively, where (dcbpyH2 is
4,4′-dicarboxy-2,2′-bipyridine; and tdbpy is 4,4′-tridecyl-2,2′-bipyridine) (Nazeeruddin et al. 2005).
Heterolepticpolpyridine ruthenium complex cis-Ru(L1)(L2)(NCS)2 (where L1 is 4,4′-dicarboxylic
acid-2,2′-bipyridine (dcbpy); and L2 is 4,4′-bis[p-diethylamino]-α-styryl]-2,2′-bipyridine) has a high
molar extinction coefficient as compared to organic dye and 8.65% conversion efficiency (Jiang et al.
2008). Novel Ru(II) phenanthroline complex [(Ru(II)(4,4′,4″-tri-tert-butyl-2,2′:6′,2″-terpyridine)-(4,
7-diphenyl-1,10-phenanthroline-disulfonicaciddisodiumsalt) (thiocyanate)] (K328) has been synthe-
sized by Erten-Ela et al. (2014).

6.6.1.2  Other Metal Complexes


A DSSC based on a TiO2 film using an I−/I3− liquid electrolyte was described by Wang et al. (2005b).
This cell produced 85% incident photon-to-current efficiency with a highly efficient (5.6%) solar
cell in the case of porphyrins-based dyes. Iron(II) complex di(aqua)bis(oxalato)iron(II), [Fe(II)
(H2O)2(C2O4)1]2−, coated with bromopyrogallol ligand has shown improved photovoltaic property
Dye-Sensitized Solar Cells 91

when compared with a bromopyrogallol ligand–coated photovoltaic device. Bromopyrogallol ligand


complexed with iron(II) shows higher stability for photodegradation and photocurrent conversion
efficiency (Jayaweera et al. 2001). Carboxylate-derivatized {CuIL2} complexes have been used in
DSSCs, and they show long-term stability and are amazingly efficient compared to ruthenium com-
plex dye (Bessho et al. 2008). Ir(III) complexes [Ir(C^N^C)(ptpy-COOH)]+ (C^N^C = 2,6-diphenyl-
pyridinato; 2,4,6-triphenylpyridinato, ptpy-COOH = 4′-(4-carboxyphenyl)-2,2′:6′,2“-terpyridine),
and [Ir(C^N^C) (tpy-COOH)]+  (C^N^C  =  2,6-diphenylpyridinato; 2,6-ditolylpyridinato, tpy-
COOH  =  4′-carboxy-2,2′:6′,2″-terpyridine) have been synthesized as photosensitizers and
compared with cis-diisothiocyanatobis(4,4′-dicarboxy-2,2′-bipyridine)ruthenium(II) (N3) photo-
sensitized DSSCs. The Ir(III) complex shows a 2.16% power conversion efficiency and a longer
lifetime (Shinpuku et al. 2011). The Pt(II)-based dye 4,4′-dicarboxy-2,2′-bipyridine and quinox-
aline-2,3-dithiolate ligands were used in DSSCs, and it gave a 2.6% power conversion efficiency
(Islam et al. 2000).
The Os complex [Os(II)H3tcterpy)(CN−)]− (where H3tcterpy is 4,4′,4″-tricarboxy-2,2′:6′,2″-
terpyridine) showed much better photochemical stability as compared to the Ru sensitizer for
DSSCs (Argazzi et al. 2004). Mordant dye has a salicylate chelating group, which is effective for
the surface of a nanocrystalline TiO2. Forty-nine types of mordant dyes have been used in DSSCs.
When mordant dye was compared with N3 dye, it showed more than 0.2 mA photocurrent, and it
bound more strongly with the TiO2 surface (Millington et al. 2007)

6.6.2  Dyes
6.6.2.1  Metal Complex Free Dyes
The Ru(II) complex dyes have shown the best results in DSSCs, but these are very costly because
of ruthenium, which may increase the overall cost of the cell. Some researchers developed cost-
efficient dyes for dye-sensitized solar cells to overcome this problem.
The DSSC sensitized with benzothiazolemerocyanine(3-carboxyalkyl-5-[2-(3-alkyl-2-benzothi-
azolinyldene)ethylidene]-2-thioxo-4-thiazolidinone dye has given 4.5% cell efficiency, which was
higher than that of other organic dyes. Dye molecules were aggregated on a TiO2 surface. The J
aggregation of the dye with long alkyl chains showed excellent capability for sensitization on the
TiO2 electrode. Power conversion efficiency and incident photon-to-current conversion efficiency
increase when alkyl chain attached to the benzothiazole ring increases. The dye molecules were
fixed by a chelate-like linkage of the carboxylate group on the TiO2 surface (Sayama et al. 2002).
The indoline dye has given the highest power conversion efficiency as compared with the metal-free
organic dyes. Using 4-tert-butyl pyridine as the electrolyte and cholic acid as a co-adsorbent, 8.00%
efficiency was obtained, whereas it gives 6.51%, 7.89%, and 8.26% power conversion efficiency with
N3 dye and N719 dye in the same conditions (Horiuchi et al. 2004). The coumarin derivatives were
found to work as excellent photosensitizers for the DSSCs. These cells performed with 5.6% power
conversion efficiency (Hara et al. 2001). Other coumarin derivatives such as thiophene have been
used in DSSCs, and 7.7% solar-to-electrical power conversion efficiency was achieved by Hara et al.
(2003). Oligothiophene-containing coumarin dyes (NKX-2677) have also been developed by Hara
et al. (2005), and 7.4% conversion efficiency of the cell was achieved. Polypyridine rings contain-
ing dyes were used as sensitizer in solar cells consisting of nanocrystalline titania, and they were
examined by in situ micro and macro Raman spectroscopy (Kontos et al. 2008).
Organic  chromophore dyes (3,4-ethyldioxythiophene and thienothiophene) were designed for
DSSCs by Zhang et al. (2009). This dye has shown 9.8% power conversion efficiency and excellent
stability of the solar cells. Some imidazole derivatives CD-4 (4.11%) and CD-6 (1.15%) triphenyl-
amine-based organic dyes were used by Chen et al. (2014b). The JK-16 and JK-17 organic dyes
containing bis-dimethylfluorenyl amino benzo[b]thiophene were synthesized and gave 7.43% and
5.49% power conversion efficiency in DSSCs, respectively (Choi et al. 2007). Chang et al. (2011)
92 Solar Energy Conversion and Storage

introduced a highly efficient porphyrin sensitizer containing two ortho-substituted long alkoxyl
chains for dye-sensitized solar cells. They achieved 10.17% conversion efficiency.
Metal-free dye was a better option than organic metal-containing dyes for further improvement
of the sensitizer. These dyes gave almost the same output for DSSCs, and with a low fabrication
cost. The metal-free organic dyes containing thienothiophene and thiophene segments were syn-
thesized for DSSCs with a 6.23% solar-to-electricity conversion efficiency (Li et al. 2006). Three
metal-free organic dyes (H11–H13) containing 3,6-disubstituted carbazole, benzothiadiazole, and
cyanoacrylic acid were used by Lee et al. (2013). The H13 dye has given the best results when
used with co-adsorbed chenodeoxycholic acid (CDCA) for sensitization of the TiO2 nanocrystal-
line surface. The cell has displayed open-circuit voltage of 0.71 V, short-circuit current density of
12.69 mA cm–2, fill factor of 0.71, and reasonably good efficiency of 6.32%.
Two thiophene-(N-aryl)pyrrole-thiophene (TPT)-based metal-free dyes have been introduced by
Tamilavan et al. (2014). The performances of solar cells with these dyes in the presence and absence
of co-adsorbent were reported as 7.06% and 6.85%, respectively. Cosensitized DSSC with metal-
free squaraine dye (SQ2) and diarylaminofluorene-based organic dye (JD1) with power conversion
efficiency 6.36% has been reported by Lin et al. (2014).

6.6.3  Natural Pigments


Ru(II) complex dyes, other organic dyes, and metal-free dyes performed very well as sensitizers
in DSSCs and gave good efficiency, but these dyes are limited in number and are costly. Natural
pigments (plant pigments) can be used as sensitizers for low-cost fabrication of dye-sensitized solar
cells. Chlorophyll, carotene, cyanin, and cyanidine are freely available in plant leaves, flowers,
and fruits (Tennakone et al. 1997). Natural pigments have the advantage of being environmentally
friendly freely available in nature.
Natural pigments extracted from rosella, blue pea, and a mixture of both extracts have been
reported for the fabrication of DSSCs by Wongcharee et al. (2007). The mixture of both extracts
adsorbed by the TiO2 surface did not show any synergistic light absorption and photosensitization
when compared with each extract alone. The power conversion efficiencies of the DSSCs were
0.37%, 0.05%, and 0.15% when sensitized with rosella extract, blue pea extract, and the mixed
extract, respectively. The efficiency of DSSCs photosensitized with rosella extract was increased
from 0.37% to 0.70%, when the aqueous solution of dye was extracted at 100°C, and the pH of
the extract was varied from 3.2 to 1.0. The extracts of natural fruits such as black rice, capsicum,
Erythrina variegata flower, Rosa xanthina, and kelp have also been used as sensitizers in the fab-
rication of DSSCs (Hao et al. 2006). The black rice extract performed as the best sensitizer due
to better interaction between the carbonyl and hydroxyl groups of anthocyanin molecule. Water-
soluble natural pigment betalain (from beet roots) has good oxidizing capacity with strong visible
light absorption. The cell was assembled with nanocrystalline TiO2 as a photoanode treated with
ethanolic HCl solution and sensitized with betanin solution. Then 2.42  mA  cm−2 photocurrent
and 0.44 V photovoltage were achieved using methoxypropionitrile containing I−/I3− redox elec-
trolyte (Zhang et al. 2008). Zhou et al. (2011) used 20 different natural dye extracts from natural
materials such as flowers, leaves, fruits, traditional Chinese medicines, and beverages. The open-
circuit voltages varied from 0.337 to 0.689 V, and short-circuit photocurrent varied from 0.14 to
2.69 mA cm−2. They achieved maximum photo-to-electric conversion efficiency from mangosteen
pericarp (1.17%).
Yusoff et al. (2014) reported the use of anthocyanin pigments from 35 native plants as potential
sensitizers for DSSC. Melastoma malabathricum (fruit pulp), Hibiscus rosa-sinensis (flower), and
Codiaeum variegatum (leaves) have given the highest absorption peaks. Melastoma malabathricum
has the highest amount of anthocyanin pigments (8.43 mg/L), and it gave the power conversion effi-
ciency of 1.16%, while with Hibiscus rosa-sinensis and Codiaeum variegatum, 0.16% and 1.08%
power conversion efficiencies have been achieved, respectively.
Dye-Sensitized Solar Cells 93

6.7 ELECTROLYTES
Electrolytes are a most effective and important component in DSSCs. The stability and conversion
efficiency of cells depend on the performance of the electrolytes. Three types of electrolytes have
been used in DSSCs: liquid electrolyte (organic solvent and ionic liquid electrolyte), quasi-solid-state
electrolyte, and solid-state electrolyte.
The basic function of a redox couple in the electrolyte is to regenerate the dye molecule after
electron injection into a conduction band of a semiconductor and to transport the positive charge
(holes) toward the counterelectrode. Mostly, a low-volatility solvent-based liquid electrolyte, I3–/I–
redox couple in acetonitrile, has been used in DSSCs.

6.7.1 Liquid Electrolytes
The iodide/triiodide redox shuttle was widely used in DSSCs with ruthenium dye (H2-dcbpy)
Ru(NCS)2 (H2-dcbpy = 2,2′-bipyridine-4,4′-dicarboxylic acid) as sensitizer, TiO2 as a working
electrode, and a platinum quasi-reference electrode. It was observed that in the presence of Li+,
the photocurrent was five times greater than when (C4H9)4N+ was used. Some other redox shut-
tles used include ferrocene, thiocyanate, triiodide, and bromide, but an iodide/triiodide redox
couple was found to be the most dominant (Wolfbauer et al. 2001). Paulsson et al. (2003) inves-
tigated trialkylsulfonium iodides–based electrolytes using liquid salts of (Et2MeS)I, (Bu2MeS)
I, and (Bu2EtS)I, where iodine was present in low concentrations. Among them, iodine-doped
(Bu2MeS)I electrolyte (3.7%) has shown the best performance for DSSCs. Other types of ionic
liquids such as 1-ethyl-3-methylimidazolium bis(trifluoromethane sulfonyl)imide (EMImTFSI),
1-ethyl-3-methylimidazolium tetrafluoroborate (EMImBF4), 1-butyl-3-methylimidazolium hexaflu-
orophosphate (BMImPF6), 1-ethyl-3-methylimidazolium dicyanamide (EMImDCA), and 1-butyl-
pyridinium bis(trifluoromethane sulfonyl)imide (BPTFSI), where 1-ethyl-3-methylimidazolium
iodide (EMImI) and I2 were dissolved as a redox couple, were used by Kawano et al. (2004). The
EMImDCA (1-ethyl-3-methylimidazolium dicyanamide) and I2 electrolyte have given 5.5% effi-
ciency in a cell, when [I−] + [I3−] = 2 M and [I−]:[I2] = 10:1 were used with 4-t-butylpyridine and
LiI. The fabrication of large (100  mm  ×  100  mm) dye-sensitized solar cells was accomplished
by Matsui et al. (2004). Liquid electrolyte containing EMIm–TFSA (1-ethyl-3-methylimidazolium
bis(trifluoromethanesulfonyl) amide) was used as the redox system. The power conversion efficiency
of 2.7% in a large-sized DSSC was achieved, while a small-sized (9 mm × 5 mm) DSSC showed
4.5% energy conversion efficiency.
Wang et al. (2003a) introduced the solvent-free ionic liquid electrolyte based on 1-methyl-3-pro-
pylimidazolium iodide, 1-methyl-3-ethylimidazolium dicyanamide, and LiI. The cell was sensitized
with an amphiphilic polypyridyl ruthenium dye. The addition of LiI increases the dye regeneration
rate and electron injection yield, and the cell performed with 6.6% power conversion efficiency at high
light irradiation, while it showed 7.1% conversion efficiency at lower light intensities. The multicol-
ored solar cell was fabricated by using nanocrystalline TiO2 electrodes sensitized with organic dyes
and platinum as a counterelectrode. A mixture of iodine and 1-butyl-3-methylimidazolium iodide in
a 0.2:10 molar ratio was used. This multicolor DSSC achieved 2.1% efficiency (Otaka et al. 2004).
The low-viscosity binary ionic liquid PMII (1-propyl-3-methyl-imidazolium iodide) and
EMIB(CN)4 (1-ethyl-3-methyl-imidazolium tetracyanoborate) electrolyte have also been used in the
DSSCs. Ru complex, Ru(2,2′-bipyridine-4,4′-dicarboxylic acid)(4,4′-bis(2-(4-tert-butyloxy-phenyl)
ethenyl)-2,2′-bipyridine) (NCS)2 has been used as the photosensitizer. Different ratios of PMII/
EMIB(CN)4 liquid electrolytes have been used, and 7.6% power conversion efficiency was obtained
(Kuang et al. 2007).
Ionic liquid electrolyte contains nitrile and vinyl function groups attached on imidazolium cations
with certain anions, for example, iodide bis[(trifluoromethyl)sulfonyl]imide ([TFSI]−) or dicyana-
mide ([N(CN)2]−) has been prepared by Mazille et al. (2006). Jovanovski et al. (2006) presented the
94 Solar Energy Conversion and Storage

use of trimethoxysilane derivative of propylmethylimidazolium iodide as gel electrolyte for DSSCs.


The overall efficiency of 3.2% was achieved when these derivatives mixed with iodine were used.
High-performance dye-sensitized solar cells with electrochemically stable nonvolatile electrolytes
were studied by Dai et al. (2006). Ethylmethylimidazolium-based ionic liquid electrolyte with dicy-
anamide anion was found to be most promising. Therefore, two different anions, tricyanomethanide
and thiocyanate, were also used. An increased conversion efficiency (2.1%) was observed in the case
of tricyanomethanide, while it was 1.7% for thiocyanate anion. The mixed solvent of tetrahydrofu-
ran (THF) and acetonitrile gave a remarkable (9.74%) conversion efficiency (Fukui et al. 2006). To
overcome the problem of stability and fabrication with ionic liquid electrolytes, some other types of
electrolytes were developed, which have good stability and are efficient in photoconversion.

6.7.2 Polymer Gel Electrolytes


Ionic liquid EMIm-TFSI (1-ethyl-3-methylimidazolium bis(trifluoromethylsulfonyl) imide) was used
as the dispersion medium for carbon nanotubes, titanium dioxide, and other carbon nanoparticles
(Usui et al. 2004). It was ground and centrifuged to give an ionic nanocomposite gel liquid. They
observed enhanced conversion efficiency as compared to DSSCs with bare ionic liquid electrolytes.
A poly(vinylidenefluoride-co-hexafluoropropylene) (PVDF-HFP)-based polymeric solid electrolyte
film was used by Asano et al. (2004) in a DSSC. The use of such a film was accompanied by a
decrease in short-circuit current density. The cell gap of the DSSC and the diffusion coefficient of
triiodide were found to play a major role in increasing this current density. It was observed that short-
circuit current density of a polymeric solid electrolyte–based DSSC with 20 µm cell gap was almost
97% to that of a liquid electrolyte–based DSSC. This type of PVDF-HFP–based gel polymeric solid
electrolyte was also used by Nishikitani et al. (2006). They observed that short-circuit current was
found to increase by either increasing the diffusion coefficient of the triiodide and/or iodide ion or
by decreasing the cell gap, or both. In general, the narrowing of the cell gap is a simple and effec-
tive way to increase short-circuit current density on account of back-electron transfer from titania to
an oxidized dye, dependent on short-circuit current density. The photon-to-electron conversion effi-
ciency of 2.93% was obtained when the semisolid ionic gel electrolyte was prepared from agarose,
natural polysaccharide, and 1-alkyl-3-methyl-imidazolium salts (Suzuki et al. 2006). A polymer gel
electrolyte, poly(vinylidenefluoride-co-hexafluoropropylene) was used by Lee et al. (2008a) in DSSC
with graphite and titania nanoparticles as fillers. Graphite nanoparticles were found to be better fill-
ers than titania. A DSSC with polymer gel electrolyte with graphite nanoparticles has cell efficiency
of 6.04% as compared to 4.69% (without filler), which was close to that of a liquid system.
Park et al. (2012) introduced a new type of UV-cured resin polymer gel with the addition of
GBL  (γ-butyrolactone). This electrolyte has long-term stability and high photo-to-electric con-
version efficiency. The iodide-conducting plastic crystal electrolytes based on N,N-dimethyl-
2-(methylsilyloxy) ethanaminium cations and I–/I3– anions was reported by Bertasi et al. (2014).
Arof et al. (2014) developed two mixed iodide salt systems: the mixture made of potassium iodide
(KI) with a small K+ cation and tetrapropylammonium iodide (Pr4NI) with a bulky Pr4N+ cation.
When KI and Pr4NI electrolytes were used alone for the cell, they gave 2.37% and 2.90% efficient
cells, respectively, whereas the mixed iodide system of KI:Pr4NI gave a 3.92% efficient cell. The
PVDF-co-HFP (poly(vinylidenefluoride-co-hexafluoropropylene):side-chain liquid crystal polymer
(SCLCP) electrolytes have been developed for use in DSSCs and exhibited 5.63% efficiency. This
electrolyte may increase the redox couple reduction and reduction charge recombination as com-
pared to PVDF-co-HFP-based polymer electrolyte (Cho et al. 2014).

6.7.3 Other Electrolytes
Wang et al. (2004b) presented solvent-free ionic electrolyte as an alternative to iodide/triiodide for
nanocrystalline solar cells. The electrolyte consists of a SeCN−/(SeCN)3− redox couple, and the cell
Dye-Sensitized Solar Cells 95

reached 7.5%–8.3% power conversion efficiency. After that, Wang et al. (2005c) introduced a more
stable Br−/Br3− redox couple in a dye-sensitized solar cell. These cells were sensitized with eosin
Y with liquid electrolyte (prepared with the combination of LiBr + Br2 in acetonitrile) and gave
2.61% power conversion efficiency, while LiI + I2 gave only 1.67%. An increase in power conver-
sion efficiency to 56% was observed when I3−/I− was replaced with Br−/Br3−. Bergeron et al. (2005)
used (SeCN)2/SeCN− or (SCN)2/SCN− or I3−/I− and mesoporous nanocrystalline SnO2 as a redox
couple pseudohalogen/halogen electrolyte and semiconductor in regenerated solar cells. They used
three sensitizers, Ru(deep)(bpy)2(PF6)2, Ru(deep)2(dpp)(PF6)2, and Ru(deep)2(bpz)2(PF6)2 (deep is
4,4′-diethylester-2,2′-bipyridine; dpp is 2,3-dipyridyl pyrazine; and bpz is bipyrazine). The current
conversion efficiencies and open-circuit voltages were found comparable for (SeCN)2/SeCN− and
I3−/I−. The binding of sensitizers (adsorption) to the SnO2 surface was described by the Langmuir
model. Sapp et al. (2002) made use of some substituted polypyridine complexes of cobalt(II/III) as
efficient electron transfer mediators in DSSC. Tris(4,4′-di-tert-butyl-2,2′-dpyridyl)cobalt(II/III) per-
chlorate was found to be the best mediator with 80% efficiency, better than that of iodide/triiodide-
based DSSCs. This ligand is simple and readily available; therefore, its complex with cobalt as the
mediator can be considered as a practical alternative for an efficient electron transfer mediator in
DSSC. It is a nonvolatile as well as noncorrosive complex. It was observed that gold and carbon are
superior cathode materials to platinum with no evidence of corrosion.
Other metal-based electrolytes like tris(bipyridine) complex of cobalt(II/III) have been used as
redox couple electrolytes (Pazoki et al. 2014). The effect of dye coverage of mesoporous titania elec-
trode was also observed using D35 dye. The adsorption of the dye on TiO2 was reasonably fit as per
the Langmuir adsorption isotherm. It was observed that the electron lifetime in the DSSC increased
remarkably with an increase in dye coverage on the TiO2 electrode. As a result, there was a dra-
matic increase in solar cell efficiency. Maruthamuthu et al. (2006) designed a lipid (detergent)-based
composite dye-sensitized solar cell without using liquid or solid electrolyte. They used a (cis-Ru(II)
(LH2)2(NCS)2)-TiO2 layer exposed to Triton X (C16H26O2)n, ion-conducting materials, and dissolved
iodine with surface-treated powder of carbon. Photovoltage (0.45 V), photocurrent (1.5 mA/cm2),
and energy conversion efficiency (0.22%) have been observed under solar light.
Recently, Luz et al. (2013) have prepared BiOCl and BiOBr nanodiscs via water-based nucle-
ation and purified these samples by a phase transfer reaction. The nanodiscs were finally deposited
as a porous p-type semiconductor layer. Coumarin 343 was used as a sensitizer with BiOCl/KI-I2
system, while eosin Y was used as sensitized with a BiOBr/[C4MPyr]2[Br20] (C4MPyr = N-butyl-N-
methylpyrrolidinium) system. They have applied such a polybromide electrolyte for the first time
in a DSSC. These systems show 40.6% and 28.6% fill factor and 0.003% and 0.0005% conversion
efficiency, respectively. The ionic liquid crystals C12EImI (1-dodecyl-3-ethylimidazolium iodide)
and C10EImI (1-decyl-3-ethylimidazolium iodide) have been synthesized by Xu et al. (2013). Both
electrolytes have the same electrochemical and thermal stability, but the I− diffusion coefficient and
photon conversion efficiency of C12EImI were higher than that of C10EImI-based DSSCs.

6.8 ELECTRODES
6.8.1  Nanocrystalline Semiconductor Working Electrode
The semiconductor is the heart of the DSSC. It is composed with nanoparticles of mesoporous oxide
thin layer. Generally in a DSSC, oxides like TiO2, ZnO, SnO2, Nb2O5, In2O3, WO3, Ta2O5, and ZrO2
are used as semiconductor materials. The TiO2 is used widely as a semiconductor due to its avail-
ability in the market, nontoxicity, low cost, environmentally friendly behavior, and biocompatibility.

6.8.1.1  Naïve Semiconductors


Nanocrystalline TiO2 particles were designed with high surface area (10–20 nm diameter) to improve
light absorption by the semiconductor material. The mesoporous morphology of thin film plays an
96 Solar Energy Conversion and Storage

important role in collecting sunlight on an oxide film. When light penetrates the dye-covered oxide
“sponge,” it crosses hundreds of adsorbed dye monolayers. Photons with energy close to the absorp-
tion maximum of the dye were absorbed completely. Hence, such a mesoporous material mimics the
process of light absorption by green leaves as in the case of photosynthesis. Different methods of
preparing TiO2 nanocrystalline semiconductors were proposed by Li et al. (1999). They introduced
the sol-gel method of preparing a TiO2 (100 nm diameter) nanoparticle thin layer with the help of
spin-coating or doctor-blading techniques. Modification in the sizes of the nanoparticles of the TiO2
thin layer (25 nm diameter) and fabrication with screen printing have resulted in maximum power
conversion efficiencies, due to the large internal surface areas of nanocrystalline particles.
The colloidal TiO2 film has been prepared in the presence of flexible substrates without using any
organic surfactant by sintering at 100°C, and its conversion efficiency was obtained as 1.2% (Pichot
et al. 2000). The comparative study of rutile and anatase TiO2-based dye-sensitized solar cells has
been presented by Park et al. (2000). The scanning electron microscopy (SEM) study shows that the
rutile films consist of homogeneously distributed rod-shaped particles with an average dimension
of 20 × 80 nm. The photovoltage shows a small change for anatase and rutile TiO2 film–based solar
cells, whereas the photocurrent of a rutile-based cell was 30% lower than that of an anatase-based
solar cell. The overall power conversion efficiency was close for both the dye-sensitized rutile and
anatase TiO2 films at the same intensity. Lindstrom et al. (2001) demonstrated a new method for
preparing a nanostructured porous layer of TiO2 film on conductive glass or plastic substrate. The
cell generated 5.2% solar-to-electrical power conversion efficiency when plastic film was used.
Miyasaka et al. (2002) presented a new nonsintering method for preparing a nanoporous semi-
conductor thin film, where TiO2 particles were electophoretically deposited on the electrode. The
nanoporous TiO2 was post-treated with the help of chemical vapor deposition of Ti alkoxide and
microwave irradiation. It shows the conversion efficiency of 4.1% under irradiation of AM 1.5 light
(100 mW cm−2). A wide-bandgap nanocrystalline TiO2 film was electrophoretically deposited with-
out using any surfactant or post-thermal treatment (Yum et al. 2005). This film was used in flex-
ible dye-sensitized solar cell on the electrode film. The fill factor and energy conversion efficiency
of such a flexible dye-sensitized solar cell was 50.0% and 1.03%, which was raised to 56.3% and
1.66%, respectively, on compression.
Ito et al. (2006) described the fabrication of a highly efficient, flexible, dye-sensitized solar cell
with a maximum 7.2% conversion efficiency in a plastic base DSSC. This cell was fabricated with
Ti metal foil substrate as the photoanode and Pt-electrodeposited as counterelectrode on ITO/PEN
(polyethylene naphthalate). Koo et al. (2008) investigated the effect of light scattering on the conver-
sion efficiency of a DSSC. They used two different types of rutile TiO2 nanoparticles with diameters
of 0.3 µm (G1) and 0.5 µm (G2). The conversion efficiency of G1 and G2 TiO2 with a 7 µm thick
main layer film was improved from 7.55% to 8.94%, but when a 14 µm thick main layer film was
used, the power conversion efficiency improved from 8.60% to 9.09% and 9.15%, respectively, upon
depositing G1 and G2 particulate overlayers.
The fabrication of dye-sensitized solar cells based on homemade ZnO nanoparticles was reported
by Lu et al. (2010), and its conversion efficiency was obtained as 3.92%. Anta et al. (2012) analyzed
the performance of a ZnO-based, dye-sensitized solar cell. ZnO has the unique combination of
high bulk electron mobility and the richest variety of nanostructures with a wide range of synthesis
routes. The comparative study of Nb2O5- and TiO2-based solar cells was reported by Sayama et
al. (1998). The cell was sensitized with ruthenium(II) cis-bis(thiocyanato)bis(2,2′-bipyridyl-4,4′-
dicarboxylic acid) complex. The Ru dye was attached to the surface of Nb2O5 by ester linkage. The
IPCE performance of the Nb2O5 photoelectrode was improved to 32% at 548 nm by the treatment of
Nb alkoxide, and the overall solar-to-electric energy conversion efficiency obtained was 2%.
A DSSC was fabricated by Kumara et al. (2006) using a highly conductive and optically trans-
parent glass consisting of an inner layer of indium tin oxide (ITO) and an outer layer of fluorine-
doped tin oxide (FTO). The nanocrystalline films of titania with large (30 nm) and small (5 nm)
sized particles can promote porosity and light scattering. It was observed that incorporation of trace
Dye-Sensitized Solar Cells 97

quantities of MgO into TiO2 increases the efficiency of the solar cell. There was an increase in the
efficiency of the solar cell from 5.6% in the absence of magnesium oxide to 7.2% in the presence of
magnesium oxide. They have also discussed the mechanism of improvement of cell performance on
incorporating magnesium oxide.

6.8.1.2  Modified Semiconductors


Dye-sensitized solar cells are made of nanocrystalline films of some oxide materials, but the depo-
sition of some ultrathin shells of insulators or high bandgap containing semiconductors on crystal-
lites can make dramatic changes (Tennakone et al. 2002). The modified TiO2 nanoparticles coated
with Mg(OH)2 were prepared by Jung et al. (2005). The specific surface area of an electrode was
prepared with the help of a nanocrystalline core–shell-structured TiO2. The cell performance was
increased to 45% as compared to uncoated TiO2 electrode. Thus, it can be concluded that the spe-
cific surface area of nanoparticles increases the conversion efficiency.
The performance of DSSC was compared before and after treatment of TiO2 electrode with an
electrochemical reaction in the presence of metals such as Mg, Zn, Al, and La in 2-propanol (Yum
et al. 2006). On coating metal hydroxides without a sintering process, the magnesium hydroxide
gave the maximum enhancement in photovoltage, improved fill factor, and overall conversion effi-
ciency of the solar cell. The overcoating of TiO2 with magnesium hydroxide functions as the block-
ing layers at the fluorine-doped tin oxide and TiO2 interfaces.
Lee et al. (2006) fabricated the working electrode with core–shell-type nanoparticles. Here, TiO2
and CaCO3 were used as cores and shells, respectively. This type of overlayering makes this cell
more efficient than an uncoated TiO2 layered electrode. It created maximum dye adsorption due to
the presence of a high isolated point of the coated layer and anticipation of the back-electron transfer
by the insulating nature of the overcoating of the layer. Three different layered TiO2 photoelectrodes
were designed and investigated by Hu et al. (2007). These layers were constructed in three different
parts like small pore–size films, large pore–size films, and light-scattering particles on conducting
glass according to required thickness. Thus, the photovoltaic performance of a cell can be affected
by different porousness, pore-size distribution, and surface area. Good power conversion efficiency
(6%) was obtained with 15 × 20 cm2 surface area in AM 1.5 sunlight (100 mW cm−2).
The Al2O3-coated nanoporous TiO2 thin-film dye-sensitized solar cell showed retardation of the
interfacial charge recombination and improved efficiency of the solar device 30% (Palomares et al.
2002), whereas an MgO-coated TiO2 electrode was prepared by the surface modification method
in the presence of DC magnetron sputtering (Wu et al. 2008). By this modification, dye adsorption
was increased, which decreases in trap states and clamp interfacial recombination. The efficiency
was increased from 6.45% to 7.57%.
Senevirathna et al. (2007) demonstrated SnO2/MgO film dye-sensitized solar cells. In this type of
cell, an SnO2 crystalline surface was covered with an ultrathin shell of MgO; this electrode protests
dye and electrolytic degradation better than TiO2-based solar cells. Li et al. (2013) introduced a one-
step electrophilic deposition method for the preparation of MgO-coated SnO2 photoanodes for dye-
sensitized solar cells. The MgO-coated SnO2 semiconductor was prepared from a facile method.
The photoanode was hydrolyzed at 413 K and irradiated under UV light. The post-treatment makes
the photoanode more flexible. The overall power conversion efficiency in this case reached 5.48%.
The fabrication of a DSSC with core–shell MgO-TiO2 electrode was done by a simple chemical
bath deposition method to coat a thin MgO film around TiO2 nanoparticles (Li et al. 2010). The
TiO2-MgO electrode showed excellent performance as compared to a noncoated TiO2 electrode.
The isc, Voc, and fill factor were found to be 8.80 mA cm−2, 646 mV, and 0.69, respectively, and power
conversion efficiency increased from 4.32% to 5.26%. The surface modification of TiO2/MnTiO3
was made by Shaterian et al. (2014). MnTiO3 nanoparticles were prepared by sol-gel method by add-
ing manganese acetate and tetrabutyltitanate as Mn and Ti sources and stearic acid as complexing
agent. The modified photoanode shows increased power conversion efficiency and improvement in
photovoltage due to increasing electron lifetime on the TiO2/MnTiO3 interface.
98 Solar Energy Conversion and Storage

6.8.1.3  Doped Semiconductors


Various methods have been developed for improving the photocatalytic activity of TiO2 and other
semiconductors in the presence of visible light, like the doping of oxide with metals, and co-doping
with metals and nonmetals.
Nitrogen-doped TiO2 electrode was used in a dye-sensitized solar cell by Ma et al. (2005). In
nitrogen-doped titania, the substitution of oxygen sites with nitrogen atom was confirmed by x-ray
photoemission spectroscopy. This powder and film showed an absorption in the visible range (400–
535 nm). Such fabricated cells with nitrogen-doped titania exhibited good stability and 8% overall
conversion efficiency. Tian et al. (2010) studied the photovoltaic performance and charge recombi-
nation in nitrogen-doped TiO2 solar cells. The flat-band potential of such a film had a negative shift,
which was attributed to the formation in the O-Ti-N bond. As a consequence, the open-circuit volt-
age improved. This cell exhibited good performance as well as stability as compared to a cell with
undoped TiO2. It was observed that the replacement of oxygen-deficient titania by nitrogen-doped
TiO2 makes the DSSC system more stable.
Mn2+- and Co2+-doped TiO2 nanoparticles were prepared using a hydrothermal method by
Shalan and Rashad (2013). The trap densities were increased by the doping of Mn and Co. The
power conversion efficiency decreased due to the high change of flat-band edge. The doping of P25,
N-TiO2, and CN-TiO2 gave 1.61%, 2.44%, and 3.31% power conversion efficiency, respectively (Sun
et al. 2013), while iodine-doped ZnO nanoparticles have shown 3.6%–4.6% cell efficiency (Zheng
et al. 2014). Electron injection and electron transportation could be controlled by introducing Cr3+
and CNTs into a TiO2 photoanode, respectively. The presence of Cr3+ improved the morphology of
the CNT-TiO2 electrode. The solar cell was composed of 3 at.% Cr3+ and 0.025 wt% CNTs, which
showed 7.47% efficiency (Massihi et al. 2013). A successive embedding of N-doping and surface
modification of TiO2 photoelectrodes with CaCO3 for dye-sensitized solar cells was reported by
Park et al. (2013). It was observed that their combined effect resulted in an increase in short-circuit
current, open-circuit voltage, as well as photoelectric conversion efficiency of the cell. The effi-
ciency of this cell was improved to 9.03% from 5.42% and 7.47% for unmodified N-doped electrode
and N-doped CaCO3 surface-modified electrode, respectively. Surface modification of an N-doped
electrode by CaCO3 formed a barrier, which suppresses the charge recombination of photogenerated
electrons to dye or the electrolyte. This extends their lifetime in the electrode and, consequently,
the cell efficiency.
Liu (2014a) reported DSSC based on Fe-doped titania electrode. Pure and Fe-doped TiO2
electrodes were prepared by the hydrothermal method. The Fe doping was done using Fe salt
[Fe(C2O4)3.5H2O]. The flat-band potential was shifted positively after Fe doping as evident from
the Mott-Schottky plot. This shift was considered the driving force for injected electrons from the
valence band of the dye to the conduction band of the semiconductor. The photovoltaic efficiency
of cell and fill factor increased from 6.07% and 0.53 to 7.46% and 0.63, respectively, as compared
to pure TiO2. Liu (2014b) also tried to improve the efficiency of the DSSC by using Mg-doped TiO2
thin films. The hydrothermal method was used for the preparation of Mg-doped titania electrode
by using Mg salt [Mg(NO3)2.6H2O]. Fluorine-doped tin oxide glass was coated with this Mg-doped
slurry and sintered at 450°C. The doped Mg ions were present in the form of Mg2+, which could act
as a trap for e− or h+, and as a result, the e−/h+ pair recombination rate is reduced. Incorporation of
Mg2+ increases the photovoltaic efficiency from 5.62% to 7.12% and short-circuit current from 14.9
to 19.1 mA as compared to undoped TiO2 thin-film electrode.
The high efficiency of a dye-sensitized solar cell depends on effective light scattering and har-
vesting, rapid electron transfer and transport, and reduced charge recombination. Wang et al. (2014a)
attempted to study the effect of the addition of ZnO into a TiO2 electrode. They observed that the
addition of only 2% zinc oxide increased dye adsorption, light scattering, and electron transfer.
This modification has enhanced conversion efficiency 9.53%, which was 12% higher than with TiO2
without the addition of ZnO.
Dye-Sensitized Solar Cells 99

A NiO photocathode–based dye-sensitized solar cell was investigated by Wang et al. (2014b).
NiO was synthesized by the microwave-assisted hydrothermal method. Lithium-doped NiO has
also been used as a photocathode, and it has shown better IPCE as well as conversion efficiency
as compared to undoped NiO photocathode. A silica layer coated with highly uniform, mono-
dispersed, hexagonal prisms was synthesized by a hydrothermal route. These NYFYE@SiO2
(NYFYE = β-NaYF4:Yb+3/Er3+) crystallites were incorporated into TiO2 porous thin film by Guo
et al. (2014). Here, with the addition of NYFYE@SiO2, improvement was seen in light scattering
and near-infrared light harvesting of the composite photoanode resulting in increased short-circuit
current and photoelectric conversion efficiency by 14% and 29%, respectively, as compared to pure
TiO2 photoanode alone. The cell performance of DSSC in this case was 7.28%.
A core–shell semiconductor was also used in DSSC by Törngren et al. (2014), like plasmonic core–
shell Au@SiO2 nanoparticles. A thin silica coating provides better stability during thermal treatment
and also against corrosive electrolyte. Stability was further assessed using temperatures up to 500°C
for thermal stability and iodide/triiodide electrolyte for chemical stability. Mercaptosilane was used as
a linker to maintain a complete silica coating on gold cores. There was almost a 10% increase in the
efficiency of DSSC on plasmon incorporation as compared to DSSC without core–shell nanoparticles.
Submicron-sized YVO4:Eu3+, Bi3+@SiO2 core–shell particles were prepared by Lai et al. (2014)
using the hydrothermal method. It was interesting to note that these particles not only enhanced
light scattering with photoanode but were also able to convert ultraviolet light to visible light. By
embedding such particles, the cell efficiency of DSSC was improved from 3.6% to 5.9%, which is
a 64% enhancement.

6.8.1.4  Other Nanomaterials


The photo-to-electric conversion efficiency of DSSC was limited due to electron–hole charge recom-
bination at the electrode–electrolyte interface. Therefore, current development in this research area
is based on alternative nanomaterials such as nanofibers, nanowires, nanorods, nanotubes, and so
on. These can be used as a photoanodes having large surface areas, low recombination rates, and
high energy conversion efficiencies.
A highly efficient dye-sensitized solar cell was fabricated by using a single-crystalline TiO2 nano-
tube as a thin film with 5% cell efficiency (Adachi et al. 2003). The ZnO nanorod, which was annealed
in N2/H2 or O2, enhanced dye adsorption due to high OH concentration on the hydrophilic surface.
After the annealing of ZnO nanorods, fill factor and cell efficiency were also increased (Chung et al.
2010). ZnO nanotubes were synthesized onto an ITO glass, which gave 1.01% cell conversion effi-
ciency (Ranjusha et al. 2011). The hybrid TiO2-coated SnO2 nanotube has been used in the fabrication
of DSSC. The cell gave 3.53% cell efficiency, and its electron recombination rate is significantly better
than that of TiO2 nanotubes, ZnO nanowires, and TiO2 nanoparticles thin films (Desai et al., 2013).
Bamboo-like shaped TiO2 nanotubes achieved 7.36% cell efficiency (Wang et al. 2014c).
A TiO2-coated SnO2 hybrid nanorod with 40 nm diameter has been synthesized by using a modi-
fied flame spray pyrolysis (FSP). The conversion efficiency was reached at 6.93%, which was much
higher than SnO2 nanorod (3.95%) and P25 (5.27%) electrodes (Huo et al. 2014). The homogeneous
composition electrode of a TiO2 macroporous material was used where TiO2 nanowire for DSSC
showed an impressive conversion efficiency of 9.51% (Wu et al. 2014). These dye-sensitized solar
cells are low in cost, environmentally friendly, and a green source of energy to meet future energy
requirements.
Lotey and Verma (2014) synthesized BiFeO3 nanowires (average diameter of 20  nm) by the
colloidal dispersion capillary force induced template-assisted technique. The synthesized nanow-
ires possess rhombohedral structures as supported by x-ray diffraction study and Fourier trans-
form infrared data. The bandgap was in the desired range (2.5 eV). A positive shift in flat-band
potential is responsible for high electron injection efficiency. The high energy conversion efficiency
of 3.02% has been achieved in this fabricated DSSC. The hybrid photoanodes containing ZnO
100 Solar Energy Conversion and Storage

nanoparticle-nanowire were fabricated by Yodyingyong et al. (2010) for use in DSSC. This hybrid
photoanode provides a direct pathway for rapid electron transport, while nanoparticles dispersed
between nanowires offer a high specific surface area for adsorption for the dye. This type of modi-
fication enhances photoconversion efficiency to a much higher percentage (4.2%) as compared to
ZnO nanoparticles (1.58%) and ZnO nanowires (1.31%).
Krysova et al. (2014) have developed a simplified method for the preparation of thick mesopo-
rous nanofiber electrodes. Four types of TiO2 nanofibers were used. These were electrospun from
polyvinylpyrrolidone (PVP) and hydroxypropylcellulose (HPC) as templates. The diameter and sur-
face area ranged between 100–280 nm and 9–100 m2/g, respectively. Mesoporous TiO2 films were
prepared by supramolecular templating with a block copolymer Pluronic P123. Then 2% or 15%
nanofibers were added to the mesoporous film. This resulted in a decrease in dye N719 adsorption
but simultaneously an increase in roughness factor. This enhances the photoelectric conversion
efficiency of the cell to 5.51%, more than that of mesoporous film (4.96%).
A hybrid photoelectrode was reported by Kong et al. (2014), which was fabricated using single-
crystalline TiO2 nanowires (NWs) inlaid with TiO2 (anatase) nanoparticles (NPs). Then 4 mm thick
vertically aligned NWs were synthesized on fluorine-doped tin oxide glass by a solvothermal route.
Nanowires provide a faster pathway for electron transport and the desired light-scattering effect.
Sidewise, inlaid NPs give an extra space for the uptake of dye. They observed 6.2% conversion
efficiency, which was about 48% more enhancement than DSSC with only NPs. This improvement
was attributed to the synergistic effect of enhanced light confinement, charge collection, and load-
ing of the dye.

6.9 COUNTERELECTRODE
The Pt electrode plays an important role in the regeneration of dye molecules by catalysis of the I−/
I3− redox couple. The Pt becomes a good choice for developing a counterelectrode due to its rough
surface area and nature toward an exposed face, but Pt is a costly noble metal, so efforts are being
made to find a low-cost alternative for production of counterelectrodes (Thomas et al. 2014).
Thus, other materials should be applied for obtaining low-cost solar cells for further development
of counterelectrodes. Graphite powder and carbon black (Kay and Gratzel 1996) and single-wall
carbon nanotubes (SWCNTs) (Suzuki et al. 2003) have been used to replace the Pt electrode. Other
carbon materials, inorganic materials, multiple compounds, polymers and composites (Wu and Ma
2014), or metal, such as platinum, gold, and nickel, and various nanostructured carbon materials,
conductive polymers (Jing et al. 2011) were also used in DSSC.
Murakami and Gratzel (2008) explained that different types of materials can be used in the
preparation of counterelectrodes like platinum, graphite, activated carbon, carbon black, single-
wall carbon nanotubes, poly(3,4-ethylenedioxythiophene) (PEDOT), polypyrrole, and polyaniline.
These may be used as catalysts for the reduction of triiodide. They prepared carbon black–loaded
stainless steel electrodes as low-cost counterelectrodes and got excellent 9.15% power conversion
efficiency. The SUS-316 stainless steel gave equivalent performance as FTO glasses.
Olsen et al. (2000) investigated the stability and performance of FTO glasses coated with vapor
deposition of Pt and Pd electrocatalyst, which were used as counterelectrodes. Saito et al. (2004) sug-
gested the porous PEDOT-TsO as a counterelectrode, which was low cost and simple to fabricate.
The potential of DSSC is further increased due to catalytic activity of the counterelectrode.
Nanosized carbon material was reported as counterelectrode with 7.56% cell efficiency (Lee et al.
2008b). Graphene nanosheets (GNs) were synthesized for DSSC with 6.81% power conversion effi-
ciency (Zhang et al. 2011). On the other hand, some nanolayers were prepared to increase catalytic
performance of the cell. PEDOT (Poly (3,4-ethylenedioxythiophene) nanoporous layers were pre-
pared by Ahmad et al. (2010). The PEDOT counterelectrode has shown 8.0% power conversion
efficiency, which is very close to that of the Pt counterelectrode (8.7%).
Dye-Sensitized Solar Cells 101

The fabrication of DSSC with graphene-PVP (polyvinyl pyrrolidone) counterelectrode coated on


FTO glasses was reported by Li et al. (2012). The PVP is linked with graphene by formation of an
ester bond (-C-O-OC-). Due to the high electrocatalytic activity of the electrode, it might show low
charge-transfer resistance and high redox current density, and it displayed 3.01% power conversion
efficiency.
Wang et al. (2013a) proposed that ZnO could be used as a counterelectrode with PEDOT:PSS
(poly(3,4-ethylenedioxythiophene):polystyrene sulfonate), and it could be an efficient counterelec-
trode with a maximum power conversion efficiency of 8.17%. Chen et al. (2013) reported an SnS
low-cost electrode for DSSC. They prepared SnS nanosheets, SnS nanowires, and SnS2 nanosheets
as alternatives to the Pt electrode. They determined power conversion efficiencies of 7.56% for Pt,
6.56% for SnS NS, 5.00% for SnS NW, and 5.14% for SnS2 NS. The remarkable performance of
SnS-based DSSC has shown that it had excellent catalytic activity for the reduction of triiodide to
iodide.
Pt-free electrodes have been synthesized by Ahmad et al. (2014). They introduced counterelec-
trodes by using graphene nanoplatelets (GNPs) or multiwall carbon nanotubes (MWCNTs). Hou
et al. (2013) introduced rust (α-Fe2O3) as a new counterelectrode, and Yue et al. (2013) prepared
PEDOT:PSS [poly(3,4-ethylenedioxythiophene):poly(styrenesulfonate)] coated counterelectrode
with high transparency and low temperature required for coating on FTO glasses. The combina-
tion of carbon nanotubes with carbon black (carbon nanocomposite) has been used as counterelec-
trode (Chen et al. 2012). Noh and Song (2014) innovated a new type of Ru/Ti bilayer on flat glass
counterelectrode, which gave 2.4% power conversion efficiency. A new type of MoS2 transparent
counterelectrode was presented by Zhang et al. (2014). To improve the catalytic activity of the elec-
trode, they created an artificial active edge site by patterning holes on MoS2 atomic layers. After
the hole patterning on MoS2 layers, DSSC has shown remarkable improvement in power conversion
efficiency from 2.0% to 5.8%. Nickel sulfide (6.81%) and cobalt sulfide (6.59%) coated FTO glasses
were also used as counterelectrodes (Yang et al. 2014). The MoSe2/Mo counterelectrode improved
the efficiency of DSSC (8.13%) as compared to Pt electrode–based DSSC (8.06%) (Chen et al.
2014a).
The rigid monolithic coal-based carbon counterelectrode has been synthesized by Wang et al.
(2014d). Coal-based carbon electrode has shown low internal resistance with 6.38% power conver-
sion efficiency. The surface modification with a mesoporous carbon catalytic layer could give 8.73%
power conversion efficiency.
A polypyrrole nickel-coated cotton fabric counterelectrode was proposed by Xu et al. (2014) with
3.3% power conversion efficiency. CuSbS2 was also used as a counterelectrode (Ramasamy et al.
2014).

6.10  OTHER TYPES OF DYE-SENSITIZED SOLAR CELL


6.10.1 Solid-State DSSC
In solid-state DSSC (SS-DSSC), redox electrolyte is replaced by a nanocrystalline solid. A p-type
semiconductor TiO2 was used because of the independent hole transport ability of the p-type semi-
conductor, which could permit charge neutralization of dye molecules after electron injection.
Dye-sensitized solar cells have achieved 11% power conversion efficiency on using liquid electro-
lytes, whereas solid-state dye-sensitized solar cells gave 5% photon-to-electric conversion efficiency
by replacing the liquid electrolyte. In order to further improve the efficiency, a variation in metal
oxides, organic and inorganic sensitizers, molecular, polymeric and electrolytic hole transporter
materials, and blocking layer was made (Yum et al. 2008). SS-DSSC is a low-cost and easily fabri-
cated alternative for amorphous silicon solar cells. It is sensitized with indoline dye, and it gave 4%
efficiency (Schmidt-Mende et al. 2005).
102 Solar Energy Conversion and Storage

Liquid electrolytes were replaced with solid-state materials like organic and inorganic p-type
semiconductors, but the power conversion efficiency of solid-state-based solar cells remained low
compared to liquid electrolyte–based DSSC. Bach et al. (1998) introduced a dye-sensitized hetero-
junction TiO2 with amorphous organic hole transport material, 2,2′,7,7′-tetrakis(N,N-di-p-methoxy-
phenyl-amine) 9,9′-spirobifluorene (OMeTAD). They got 33% yield of photon-to-current conversion
efficiency. The fabrication of highly efficient SS-DSSC with porous dye-filled TiO2 layer with ZnO
coating and molten salt capped CuI crystals was done by Meng et al. (2003). The cell achieved long-
term stability and 3.8% efficiency.
Dye-sensitized solid-state solar cells have shown low energy due to a high recombination rate at
the n-type semiconductor/dye/p-type semiconductor interfaces, when n-type semiconductor TiO2
and p-type semiconductor CuI were used. The modification of n-TiO2 crystallites with the deposi-
tion of an ultrathin layer of MgO could improve power conversion efficiency (Kumara et al. 2004).
Karthikeyan and Thelakkat (2008) reported that the pore size is more important than a large
surface area for filling the mesoporous layer with a solid-state hole conductor. The heteroleptic
Ru(II) complexes [N,N′-bis(phenyl)-N,N′-bis(3-methylphenyl)-1,1′-biphenyl-4,4′-diamine (TPD)]
have shown better efficiency due to their easy filling into pores of nc-TiO2 layer. Recently, novel
hole conductors have been synthesized with spiro-bifluorene-triphenylamine core for transporting
holes and tetraethylene glycol side chain for binding lithium ion for improvement in SS-DSSC. The
addition of Li salt could improve the performance of SS-DSC, because Li+ salt is required for the
interfacing of TiO2 and dye.
Abrusci et al. (2011) replaced spiro-OMeTAD with poly(3-hexylthiophene) (P3HT) light-
adsorbing polymer with the addition of lithium bis(trifluoromethylsulfonyl) imide salts (Li-TFSI)
in SS-DSSC sensitized with indolene-based D102. Burschka et al. (2011) doped spiro-OMeTAD
with p-type Co(III). It worked as a hole conductor for SS-DSSC and has good power conversion
efficiency (7.2%). Solid-state dye-sensitized solar cells assembled with CsSnI2.95F0.05 doped with
SnF2, nanoporous TiO2, and sensitized with N719 dye gave excellent power conversion efficiency
(10.2%). CsSnI was used for hole conduction in place of a liquid electrolyte by Chung et al. (2012).
Michaleviciute et al. (2014) synthesized a star-shaped carbazole molecule, TMPCA [tri(9-
(methoxyphenyl) carbazol-3-yl) amine] and fabricated the solid-state dye-sensitized solar cells
based on ITO/TiO2/D102/T4MPCA/Au. This cell has shown 2.23% power conversion efficiency.
Chiang et al. (2013) prepared a new tandem solid-state dye-sensitized solar cell with 3.3% power
conversion efficiency. Roh et al. (2014) used a pine tree–like TiO2 nanotube (PTT) and sensitized it
with N719 for fabricating SS-DSSC. They got an outstanding energy conversion efficiency of 8.0%
at 100 mW cm−2.
Solid electrolyte prepared with 2-hydroxyethyltrimethylammonium iodide (choline iodide) was
used in a solid-state, dye-sensitized solar cell. This solid electrolyte has a strong quaternary ammo-
nium cation, 2-hydroxyethyltrimethylammonium (HETA)+ as a cation donor. In this cell, 2.5% over-
all power conversion efficiency was achieved (Wang et al. 2013b).
Solid-state dye-sensitized solar cells fabricated with hyperbranched nanostructured TiO2 on
FTO glass and sensitized with D102 dye were described by Passoni et al. (2013). This device has
shown the highest 3.96% efficiency and 66% increased efficiency with respect to a reference meso-
porous photoanode. Fabrication of SS-DSSC with polymer substrates was done by Xue et al. (2014)
with 1.93% efficiency.
Akhtar et al. (2013) prepared a solid electrolyte with the combination of graphene (Gra) and
polyethylene oxide (PEO). Gra–PEO composite electrolyte has shown the large-scale generation of
iodide ions in a redox couple and reduced its velocity by adding LiI and I2. They achieved a 5.23%
overall power conversion efficiency. Fabrication of a SS-DSSC with P3HT [poly(3-hexilthiophene)]
as the hole transport material for the dye regeneration process with 2% power conversion efficiency
was reported by Matteocci et al. (2014).
Li et al. (2015) synthesized I−/I3−-doped 3-hydroxypropionitrile/polyaniline (HPN/PANi)
solid electrolytes for the fabrication of SS-DSSC. The I3− reduction reaction from the electrolyte/
Dye-Sensitized Solar Cells 103

counterelectrode interface to solid electrolyte system was enhanced, and the charge diffusion path
length was decreased. The HPN/PANi SS-DSSC shows 3.70% cell efficiency higher than the primi-
tive HPN-based DSSC.

6.10.2  Quasi-Solid-State DSSC


A DSSC has achieved 7% efficiency with liquid electrolyte. But liquid electrolyte creates some techni-
cal problems related to long-term stability, and the voltaic nature of organic solvents does not allow tight
sealing of the cell. From this point of view of long-term stability of the cell, liquid electrolyte could be
replaced with quasi/full solid-state electrolyte for better performance. This type of solid electrolyte was
developed and gave better long-term stability and higher efficiency than a liquid electrolyte.
A quasi-solid-state electrolyte has been used as a substitute for liquid electrolyte in quasi-solid-state
dye-sensitized solar cell (QSS-DSSC). Kubo et al. (2002) fabricated a dye-sensitized solar cell with
1-hexyl-3-methylimidazolium iodide, iodine, and low-molecular-weight gelator as a quasi-solid-state
electrolyte. They achieved 5.0% power conversion efficiency and high-temperature stability. Wang
et al. (2002) used ionic liquid polymer gel electrolyte for QSS-DSSC with 5.3% power conversion
efficiency. The polymer gel electrolyte consists of 1-methyl-3-propylimidazolium iodide (MPII) and
poly(vinylidenefluoride-co-hexafluoropropylene) (PVDF–HFP). Ionic liquid–based quasi-solid-state
electrolyte has been made from solidification of silica nanoparticles and was sensitized with ruthe-
nium polypyridyl dye. The cell gave 7% power conversion efficiency (Wang et al. 2003b).
Mohmeyer et al. (2006) prepared a stable quasi-solid-state electrolyte for QSS-DSSC with 6.3%
efficient cell by using organogelator. Cross-linked gelators containing gel electrolyte have been used
in QSS-DSSC. This electrolyte consists of ionic liquids: (1-methyl-3-propylimidazolium iodide)
solidified with polyvinyl pyridine and 1,2,4,5-tetra(bromomethyl) benzene). The solidification pro-
cess occurred after injecting the gel electrolyte into the cell (Sakaguchi et al. 2004). A new type of
ionic polymer electrolyte has been synthesized by copolymerization of alkyl-bis(imidazole)s poly-
mer, and diiodo alkyls were used in quasi-solid-state electrolyte in QSS-DSSC (Suzuki et al. 2004).
Komiya et al. (2004) introduced polymer electrolyte, which has only 7% concentration of polymer.
This polymer electrolyte was used in quasi-solid-state DSSC, and it gave 8.1% power conversion
efficiency. Poly(methyl acrylate)/poly(ethylene glycol)-based polymer gel electrolyte with 5.76%
efficiency has been reported by Shi et al. (2009), while Joseph et al. (2006) reported hybrid gel
electrolyte based on tetraethyl orthosilicate (TEOS) and poly(ethylene glycol) with 4.1% efficiency.
1-Butyl-3-methylimidazolium iodide, polyvinyl pyrrolidone, potassium iodide, and
iodine-containing ionic liquid electrolyte have been used in quasi-solid-state DSSC (5.41%) (Fan
et al. 2010). Poly(HEMA/GR) [poly(hydroxyethyl methacrylate/glycerin)]/PANi (polyaniline) gel
electrolyte has also been used for dye-sensitized solar cell. The combination of poly(HEMA/GR)
and PANi creates low charge-transfer resistance and higher electrolytic activity for the I3−/I− redox
reaction, and it gave 6.63% power conversion efficiency (Li et al. 2014). Block copolymer electro-
lyte based on PSn-b-PEOm-b-PSn (PS = polystyrene, PEO = poly(ethylene oxide) as solid hosts for
iodine/iodide was used with 6.7% efficiency in DSSC by Manfredi et al. (2014). Dkhissi et al. (2014)
reported poly(vinylidenefluoride-co-hexafluoropropylene)–based gel electrolytes for QS-DSSC
with 6.4% cell efficiency.
Fan et al. (2014) introduced metal–organic skeleton-based gel electrolytes for high-efficiency
quasi-solid-state dye-sensitized solar cell (QSS-DSSC). Metal–organic gel was prepared by coor-
dination of Al3+ and 1,3,5-benzenetricarboxylate (H3BTC). The cell fabricated with gel electrolyte
showed 8.60%, which is lower than the liquid electrolyte–based solar cell.
The novel polymeric ionic liquid gel electrolyte was synthesized from poly(oxyethylene)-
imide-imidazole complex coupled with iodide anions (POEI-II) for quasi-solid-state dye-sensitized
solar (Chang et al. 2014). POEI-II simultaneously acts as a redox mediator in the electrolyte and as a
polymer for the gelation of an organic solvent-based electrolyte, while cell durability was improved.
The QSS-DSSC with the POEI-II gel electrolyte reaches a high cell efficiency of 7.19%, while with
the POEI-II/MWCNT gel electrolyte, 7.65% power conversion efficiency was obtained.
104 Solar Energy Conversion and Storage

Shi et al. (2015) designed quasi-solid-state dye-sensitized solar cells with three-dimensional net-
work consisting poly(adipic acid pentaerythritol ester) (PAAPE). The PAAPE was prepared by the
esterification of pentaerythritol and adipic acid. This three-dimensional network gives highly effi-
cient ion-transporting channels for iodide/triiodide (I−/I3−). This gel electrolyte shows 4.03 mS cm−1
conductivity at 25°C, and the solar cell shows 6.81% power conversion efficiency. Many types of gel
polymer electrolytes were used in DSSC, including I−/I3− redox couple-contain, ethylene carbonate
(EC) and propylene carbonate (PC)-plasticized, polyacrylonitrile (PAN)-based gel polymer electro-
lyte. Jayaweera et al. (2015) reported four different types of DSSC: (1) only plasticized gel polymer
electrolyte was used in between two electrodes of the cell; (2) the same electrolyte was used, but the
gel electrolyte was deeply penetrated in the matrix of the nanocrystalline TiO2 electrode; (3) an I−/
I3− redox couple electrolyte was used with sealing the pores by PAN gel electrolyte; and (4) an I−/
I3− redox couple of liquid electrolyte was used. These cells show power conversion efficiencies for
(1) 4.1%, (2) 5.2%, (3) 8.4%, and (4) 9.8%, respectively.

6.10.3  Quantum Dot DSSC


Quantum dot dye-sensitized solar cells (QD-DSSC) have received tremendous attention as one of
the potential low-cost alternatives for p-n junction silicon solar cells. The basic difference between
a QD-DSSC and a DSSC is the material used to harvest the visible portion of the solar spectrum.
A quantum dot cell is designed with electron conductor/QD monolayer/hole conductor junctions
with high optical absorbance. Nanocrystalline semiconductor and redox electrolyte, solid-state hole
conductors were borrowed from standard DSSCs (Rühle et al. 2010).
Cadmium chalcogenide (CdX,  X  =  S, Se, or Te)-based QD-DSSC has given the best perfor-
mance, but its performance could not compete with the DSSC. The overall performance of a cell
depends on the fabrication method of the quantum dots, morphology of the photoanode, type of
electrolyte, and choice of counterelectrode (Jun et al. 2013).
The ZnO nanorods coated with TiO2 nanosheets were designed for cosensitized CdS/CdSe quan-
tum dot solar cells. TiO2 nanosheets improved the surface area of the ZnO nanorod and allowed more
adsorption for quantum dots, which gave a high short current density, an energy barrier for electron
hindrance in ZnO from being returned to the electrolyte, and reduced rate of charge recombination.
The power conversion efficiency obtained was 2.7% (Tian et al. 2013). The hybrid GaAs-based solar
cell with colloidal CdS quantum dots was demonstrated by Lin et al. (2012). CdS quantum dots have
antireflective features at long wavelengths and down-conversion in a UV regime, so they can affect
overall power conversion efficiency as compared with a traditional GaAs-based device.
Sarkar et al. (2012) reported different-sized 3-mercaptopropionic acid (MPA) stabilized CdTe quan-
tum dots (QDs) for the fabrication of QDs in ZnO nanorod (NR)-based DSSCs. They highlight two major
pathways: (1) a direct injection of charge carriers from QDs to ZnO semiconductor via photoinduced
electron transfer (PET) and (2) an indirect excitation of sensitized dye (N719) molecules by funneling
harvested light via Förster resonance energy transfer (FRET). The QD-assembled DSSCs shows higher
photoconductivity and high short-circuit density as compared to DSSC fabricated with N719 sensitizer.
The heterostructured cerium oxide quantum dots (CeO2 QDs) equipped with zinc oxide nanorods
(ZnO NRs) have been used in DSSC (Rai et al. 2014). The CeO2 QDs fitted with zinc oxide nanorods
were prepared by the combination of solvothermal and chemical bath deposition methods. The
solar cell assembled with CeO2  QDs/ZnO NRs shows improved open-circuit voltage, fill factor,
and maximum power conversion efficiency (2.65%). The CeO2  QDs/ZnO NRs shows high light-
harvesting efficiency due to the formation of a barrier layer and hindrance in back-electron transfer
on the surface of the photoanode.
Rho et al. (2014) reported silica-coated quantum dot–embedded silica nanoparticles for the fab-
rication of DSSC. This solar cell was compared with unmodified DSSC. The power conversion
efficiency was improved from 3.92% to 4.82% for SiO2/QD/SiO2 NPs DSSCs.
Fang et al. (2014) introduced graphene quantum dots (GQDs) in the fabrication of dye-sensitized
solar cells. It was observed that the amount of dye adsorption first decreased and then increased
Dye-Sensitized Solar Cells 105

with an increasing amount of GQDs in the photoanodes, while the photovoltage, photocurrent, and
conversion efficiency of these solar cells were first increased and then decreased. The role of GQDs
not only improves the properties of DSSC, but it also reduces the use of dyes. It has significant
importance in the fabrication of environmentally friendly and low-cost DSSCs.
Graphene quantum dots have also been used as cosensitizers in hybrid dye-sensitized solar cells
(Mihalache et al. 2015). GQDs were synthesized by a hydrothermal method, which permits the tun-
ing of electronic levels and optical properties by using suitable conditions for synthesis and proper
precursors. The solar cell was assembled by using TiO2/GQD/N3 Ru dye, where GQDs were used as
cosensitizer together with N3 dye. The power conversion efficiency was improved due to the energy
transfer from GQDs to N3 Ru dye by the overlapping between GQD photoluminescence and N3
Ru-dye absorption spectra. The inhibition of back-electron transfer to the electrolyte by the GQDs
improves the reduction of electron recombination to the redox couple.

6.11  APPLICATIONS OF DSSC


A DSSC is more effectively used in the field of energy harvesting because of its capability to oper-
ate under low light irradiation. Almost half of the sunlight spectrum belongs to the infrared region.
The silicon cell and CLGC solar cells adsorb sunlight at over 1000 nm, but DSSC adsorbs light at
775 nm for N719, red dye, and 900 nm for N749, black dye.
The common silicon-based solar cell can be replaced by a DSSC because of its good perfor-
mance and high efficiency (10%). In 2010, DSSCs have been fabricated with efficiency of almost
10%; hence, the commercialized DSSC modules are now attainable.
In the field of BIPV (building integrated photovoltaics), DSSCs are widely applicable in power
window and shingles or pebbles because of their multicolor options (depending on the dye). The
first building has been equipped with a wall of electric power–producing glass tiles for large-scale
testing of DSSCs. The electric power–producing glass tiles were also manufactured. DSSC can also
be used for room and outdoor calculators, gadgets, and mobile devices due to their lightweight flex-
ibility. DSSCs can be designed for indoor decorative elements due to different-colored dyes. They
can be used in many portable devices, including baggage, gears, and outfits.

REFERENCES
Abrusci, A., R. S. S. Kumar, M. Al-Hashimi, M. Heeney, A. Petrozza, and H. J. Snaith. 2011. Influence of ion
induced local coulomb field and polarity on charge generation and efficiency in poly(3-hexylthiophene)-
based solid-state dye-sensitized solar cells. Adv. Funct. Mater. 21: 2571–2579.
Adachi, M., Y. Murata, I. Okada, and S. Yoshikawa. 2003. Formation of titania nanotubes and applications for
dye-sensitized solar cells. Electrochem. Soc. 150: G488–G493.
Adams, W. G., and R. E. Day. 1876. The action of light on selenium. Proc. Roy. Soc. Lond. A 25: 113–117.
Ahmad, I., J. E. McCarthya, M. Barib, and Y. K. Gun’koa. 2014. Carbon nanomaterial based counter elec-
trodes for dye-sensitized solar cells. Solar Energy. 102: 152–161.
Ahmad, S., J.-H. Yum, Z. Xianxi, M. Grätzel, H.-J. Butt, and M. K. Nazeeruddin. 2010. Dye-sensitized solar
cells based on poly (3,4-ethylenedioxythiophene)counter electrode derived from ionic liquids. J. Mater.
Chem. 20: 1654–1658.
Akhtar, M. S., S. J. Kwon, F. J. Stadler, and O. B. Yang. 2013. High efficiency solid state dye sensitized solar
cells with graphene–polyethylene oxide composite electrolytes. Nanoscale. 5: 5403–5411.
Anderson, S., E. C. Constable, M. P. Dare-Edwards, J. B. Goodenough, A. Hamnett, K. R. Seddon, et al. 1979.
Chemical modification of a titanium (IV) oxide electrode to give stable dye sensitization without a super
sensitiser. Nature. 280: 571–573.
Anta, J. A., E. Guillén, and R. Tena-Zaera. 2012. ZnO-based dye-sensitized solar cells. J. Phys. Chem. C. 116:
11413–11425.
Argazzi, R., G. Larramona, C. Contado, and C.-A. Bignozzi. 2004. Preparation and photoelectrochemical
characterization of a red sensitive osmium complex containing 4,4′,4′′-tricarboxy-2,2′:6′,2′′-terpyridine
and cyanide ligands. J. Photochem. Photobiol. A: Chem. 164: 15–21.
106 Solar Energy Conversion and Storage

Arof, A. K., M. F. Aziz, M. M. Noor, M. A. Careem, L. R. A. K. Bandara, C. A. Thotawatthage, et al. 2014.


Efficiency enhancement by mixed cation effect in dye-sensitized solar cells with a PVdF based gel poly-
mer electrolyte. Int. J. Hydrogen Energy. 39: 2929–2935.
Asano, T., T. Kubo, and Y. Nishikitani. 2004. Electrochemical properties of dye-sensitized solar cells fabri-
cated with PVDF-type polymeric solid electrolytes. J. Photochem. Photobiol. A: Chem. 164: 111–115.
Bach, U., D. Lupo, P. Comte, J. E. Moser, F. Weissörtel, J. Salbeck, et al. 1998. Solid-state dye-sensitized mes-
oporous TiO2 solar cells with high photon-to-electron conversion efficiencies. Nature. 395: 583–585.
Becquerel, A. E. 1839. Memory on the electrical effects produced under the influence of radiation solarire.
C. R. Acad. Sci. Paris. 9: 561–567.
Bergeron, B. V., A. Marton, G. Oskam, and G. J. Meyer. 2005. Dye-sensitized SnO2 electrodes with iodide and
pseudohalide redox mediators. J. Phys. Chem. B. 100: 937–943.
Bertasi, F., E. Negro, K. Vezzu, and V. D. Noto. 2014. Iodide-conducting plastic crystals based on N,N-
dimethyl-2-(methylsilyloxy) ethanaminium cations (MESEAn+) for application in dye-sensitized solar
cells. Int. J. Hydrogen Energy. 39: 2896–2903.
Bessho, T., E. C. Constable, M. Graetzel, A. H. Redondo, C. E. Housecroft, W. Kylberg, et al. 2008. An ele-
ment of surprise-efficient copper-functionalized dye-sensitized solar cells. Chem. Commun. 3717–3719.
Burschka, J., A. Dualeh, F. Kessler, E. Baranoff, N. L. Cevey-Ha, C. Yi, et al. 2011. Tri(2-(1H-pyrazol-1-yl)
cobalt (III) as p-type dopant for organic semiconductors and its application in highly efficient solid-state
dye-sensitized solar cells. J. Am. Chem. Soc. 133: 18042–18045.
Chang, L.-Y., C.-P. Lee, C.-T. Li, M.-H. Yeh, K.-C. Ho, and J.-J. Lin. 2014. Synthesis of a novel amphiphilic
polymeric ionic liquid and its application in quasi-solid-state dye-sensitized solar cells. J. Mater. Chem.
A. 2: 20814–20822.
Chang, Y.-C., C.-L. Wang, T.-Y. Pan, S.-H. Hong, C.-M. Lan, H.-H. Kuo, et al. 2011. A strategy to design
highly efficient porphyrin sensitizers for dye-sensitized solar cells. Chem. Commun. 47: 8910–8912.
Chen, H., Y. Xie, H. Cui, W. Zhao, X. Zhu, Y. Wang, et al. 2014a. In situ growth of a MoSe2/Mo counter elec-
trode for high efficiency dye-sensitized solar cells. Chem. Commun. 50: 4475–4477.
Chen, X. Y., Y. Hou, B. Zhang, X. H. Yang, and H. G. Yang. 2013. Low-cost SnSx counter electrodes for dye-
sensitized solar cells. Chem. Commun. 49: 5793–5795.
Chen, X., C. Jia, Z. Wan, and X. Yao. 2014b. Organic dyes with imidazole derivatives as auxiliary donors for
dye-sensitized solar cells: Experimental and theoretical investigation. Dyes Pigments 104: 48–56.
Chen, Y., H. Zhang, Y. Chen, and J. Lin. 2012. Study on carbon nanocomposite counter electrode for dye-
sensitized solar cells. J. Nanomater. doi:10.1155/2012/601736
Chiang, Y.-F., R.-T. Chen, A. Burke, U. Bach, P. Chen, and T.-F. Guo. 2013. Non-color distortion for visible
light transmitted tandem solid state dye-sensitized solar cells. Renew. Energy. 59: 136–140.
Cho, W., J. W. Lee, Y.-S. Gal, M.-R. Kim, and S. H. Jin. 2014. Improved power conversion efficiency of
dye-sensitized solar cells using side chain liquid crystal polymer embedded in polymer electrolytes.
Mater. Chem. Phys. 143: 904–907.
Choi, H., J. K. Lee, K. Song, S. O. Kang, and J. Ko. 2007. Novel organic dyes containing bis-dimethylfluorenyl
amino benzo[b]thiophene for highly efficient dye-sensitized solar cell. Tetrahedron. 63: 3115–3121.
Chung, I., B. Lee, J. He, R. P. H. Chang, and M. G. Kanatzidis. 2012. All-solid-state dye-sensitized solar cells
with high efficiency. Nature. 485: 486–489.
Chung, J., J. Lee, and S. Lim. 2010. Annealing effects of ZnO nanorods on dye-sensitized solar cell efficiency.
Physica B. 405: 2593–2598.
Dai, Q., D. B. Menzies, D. R. MacFarlane, S. R. Batten, S. Forsyth, L. Spiccia, et al. 2006. Dye-sensitized
nanocrystalline solar cells incorporating ethylmethylimidazolium-based ionic liquid electrolytes.
C. R. Chim. 9: 617–621.
Dare-Edwards, M. P., J. B. Goodenough, A. Hamnett, K. R. Seddon, and R. D. Wright. 1980. Sensitisation
of semiconducting electrodes with ruthenium-based dyes. Faraday Discuss. Chem. Soc. 70: 285–298.
Desai, U. V., C. Xu, J. Wu, and D. Gao. 2013. Hybrid TiO2–SnO2 nanotube arrays for dye-sensitized solar cells.
J. Phys. Chem. C. 117: 3232–3239.
Dkhissi, Y., F. Huang, Y.-B. Cheng, and R. A. Caruso. 2014. Quasi-solid-state dye-sensitized solar cells on
plastic substrates. J. Phys. Chem. C. 118: 16366–16374.
Erten-Ela, S., S. Sogut, and K. Ocakoglu. 2014. Synthesis of novel ruthenium II phenanthroline complex and
its application to TiO2 and ZnO nanoparticles on the electrode of dye sensitized solar cells. Mater. Sci.
Semicon. Proc. 23: 159–166.
Fan, J., L. Li, H.-S. Rao, Q.-L. Yang, J. Zhang, H.-Y. Chen, et al. 2014. A novel metal-organic gel based elec-
trolyte for efficient quasi-solid-state dye-sensitized solar cells. J. Mater. Chem. A. 2: 15406–15413.
Dye-Sensitized Solar Cells 107

Fan, L., S. Kang, J. Wu, S. Hao, Z. Lan, and J. Lin. 2010. Quasi-solid state dye-sensitized solar cells based on
polyvinylpyrrolidone with ionic liquid. Energy Sources. 32: 1556–1568.
Fang, X., M. Li, K. Guo, J. Li, M. Pan, L. Bai, M. Luoshan, and X. Zhao. 2014. Graphene quantum dots opti-
mization of dye-sensitized solar cells. Electrochim. Acta. 137: 634–638.
Fukui, A., R. Komiya, R. Yamanaka, A. Islam, and L. Han. 2006. Effect of a redox electrolyte in mixed sol-
vents on the photovoltaic performance of a dye-sensitized solar cell, Solar Energy Mater. Solar Cells.
90: 649–658.
Gerischer, H., M. E. Michel-Beyerle, F. Rebentrost, and H. Tributsch. 1968. Sensitization of charge injection
into semiconductors with large band gap. Electrochim. Acta. 13: 1509–1515.
Gratzel, M. 2003. Dye-sensitized solar cells. J. Photochem. Photobiol. C: Rev. 4: 145–153.
Gratzel, M. 2005. Solar energy conversion by dye-sensitized photovoltaic cells. Inorg. Chem. 44: 6841–6851.
Guo, K., M. Li, X. Fang, M. Luoshan, L. Bai, and X. Zhao. 2014. Performance enhancement in dye-sensitized
solar cells by utilization of a bifunctional layer consisting of core-shell β-NaYF4:Er3+/Yb3+@SiO2 sub-
micron hexagonal prisms J. Power Sources. 249: 72–78.
Hao, S., J. Wu, Y. Huang, and J. Lin. 2006. Natural dyes as photosensitizers for dye-sensitized solar cell.
Solar Energy. 80: 209–214.
Hara, K., M. Kurashige, Y. Dan-Oh, C. Kasada, A. Shinpo, S. Suga, K. Sayama, and H. Arakawa. 2003.
Design of new coumarin dyes having thiophene moieties for highly efficient organic-dye-sensitized
solar cells. New J. Chem. 27: 783–785.
Hara, K., K. Sayama, Y. Ohga, A. Shinpo, S. Suga, and H. Arakawa. 2001. A coumarin-derivative dye sen-
sitized nanocrystalline TiO2 solar cell having a high solar-energy conversion efficiency up to 5.6%.
Chem. Commun. 6: 569–570.
Hara, K., Z.-S. Wang, T. Sato, A. Furube, R. Katoh, H. Sugihara, Y. Dan-oh, C. Kasada, A. Shinpo, and S.
Suga. 2005. Oligothiophene-containing coumarin dyes for efficient dye-sensitized solar cells. J. Phys.
Chem. B. 109: 15476–15482.
Horiuchi, T., H. Miura, K. Sumioka, and S. Uchida. 2004. High efficiency of dye-sensitized solar cells based
on metal-free indoline dyes. J. Am. Chem. Soc. 126: 12218–12219.
Hou, Y., D. Wang, X. H. Yang, W. Q. Fang, B. Zhang, H. F. Wang, et al. 2013. Rational screening low-cost
counter electrodes for dye-sensitized solar cells. Nat. Commun. 4: 1–8.
Hu, L., S. Dai, J. Weng, S. Xiao, Y. Sui, Y. Huang, et al. 2007. Microstructure design of nanoporous TiO2
photoelectrodes for dye-sensitized solar cell modules. J. Phys. Chem. B. 111: 358–362.
Huo, J., Y. Hu, H. Jiang, W. Huang, and C. Li. 2014. SnO2 nanorod@TiO2 hybrid material for dye-sensitized
solar cells. J. Mater. Chem. A. 2: 8266–8272.
Imahori, H., S. Hayashi, T. Umeyama, S. Eu, A. Oguro, S. Kang, Y. Matano, T. Shishido, S. Ngamsinlapasathian,
and S. Yoshikawa. 2006. Comparison of electrode structures and photovoltaic properties of porphy-
rin-sensitized solar cells with TiO2 and Nb, Ge, Zr-added TiO2 composite electrodes. Langmuir. 22:
11405–11411.
Imahori, H., T. Umeyama, and S. Ito. 2009. Large π-aromatic molecules as potential sensitizers for highly
efficient dye-sensitized solar cells. Acc. Chem. Res. 42: 1809–1818.
Islam, A., H. Sugihara, K. Hara, L.-P. Singh, R. Katoh, M. Yanagida, Y. Takahashi, S. Murata, and H. Arakawa.
2000. New platinum(II) polypyridyl photosensitizers for TiO2 solar cells. New J. Chem. 24: 343–345.
Ito, S., N.-L. C. Ha, G. Rothenberger, P. Liska, P. Comte, S. M. Zakeeruddin, P. Péchy, M.-K. Nazeeruddin,
and M. Grätzel. 2006. High-efficiency (7.2%) flexible dye-sensitized solar cells with Ti-metal substrate
for nanocrystalline-TiO2 photoanode. Chem. Commun. 38: 4004–4006.
Jayaweera, E. N., C. S. K. Ranasinghe, G. R. A. Kumara, W. M. N. M. B. Wanninayake, K. G. C. Senarathne,
et al. 2015. Novel method to improve performance of dye-sensitized solar cells based on quasi-solid gel-
polymer electrolytes. Electrochim. Acta. 152: 360–367.
Jayaweera, P. M., S. S. Palayangoda, and K. Tennakone. 2001. Nanoporous TiO2 solar cells sensitized with
iron(II) complexes of bromopyrogallol red ligand. J. Photochem. Photobiol. A: Chem. 140: 173–177.
Jiang, K.-J., J.-B. Xia, N. Masaki, S. Noda, and S. Yanagida. 2008. Efficient sensitization of nanocrystalline
TiO2 films with high molar extinction coefficient ruthenium complex. Inorg. Chim. Acta. 361: 783–785.
Jing, L., S. Ming-Xuan, Z. Xiao-Yan, and C. Xiao-Li. 2011. Counter electrodes for dye-sensitized solar cells.
Acta Phys. Chim. Sin. 27: 2255–2268.
Joseph, J., K. M. Son, R. Vittal, W. Lee, and K.-J. Kim. 2006. Quasi-solid-state dye-sensitized solar cells with
siloxane poly(ethylene glycol) hybrid gel electrolyte. Semicond. Sci. Technol. 21: 697.
Jovanovski, V., E. Stathatos, B. Orel, and P. Lianos. 2006. Dye-sensitized solar cells with electrolyte based on
a trimethoxysilane-derivatized ionic liquid. Thin Solid Films. 511–512: 634–637.
108 Solar Energy Conversion and Storage

Jun, H. K., M. A. Careem, and A. K. Arof. 2013. Quantum dot-sensitized solar cells-perspective and recent devel-
opments: A review of Cd chalcogenide quantum dots as sensitizers. Renew. Sust. Energy Rev. 22: 148–167.
Jung, H. S., J.-K. Lee, and M. Nastasi. 2005. Preparation of nanoporous MgO-coated TiO2 nanoparticles and
their application to the electrode of dye-sensitized solar cells. Langmuir. 21: 10332–10335.
Kalyanasundaram, K., and M. Grätzel. 1998. Application of functionalized transition metal complexes in
photonic and optoelectronic devices. Coord. Chem. Rev. 177: 347–414.
Karthikeyan, C. S., and M. Thelakkat. 2008. Key aspects of individual layers in solid-state dye-sensitized
solar cells and novel concepts to improve their performance. Inorg. Chim. Acta. 361: 635–655.
Kawano, R., H. Matsui, C. Matsuyama, A. Sato, M. A. B. H. Susan, N. Tanabe, and M. Watanabe. 2004. High
performance dye-sensitized solar cells using ionic liquids as their electrolytes. J. Photochem. Photobiol.
A: Chem. 164: 387–392.
Kay, A., and M. Grätzel. 1996. Low cost photovoltaic modules based on dye sensitized nanocrystalline tita-
nium dioxide and carbon powder. Solar Energy Mater. Solar Cells. 44: 99–117.
Kohle, O., S. Ruile, and M. Grätzel. 1996. Ruthenium(II) charge-transfer sensitizers containing 4,4′-dicarboxy-
2,2′-bipyridine. Synthesis, properties and bonding mode of coordinated thio- and selenocyanates.
Inorg. Chem. 35: 4779–4787.
Komiya, R., L. Han, R. Yamanaka, A. Islam, and T. Mitate. 2004. Highly efficient quasi-solid-state dye-
sensitized solar cell with ion conducting polymer electrolyte. J. Photochem. Photobiol. A: Chem. 164:
123–127.
Kong, E.-H., Y.-H. Yoon, Y.-J. Chang, and H. M. Jang. 2014. Hybrid photoelectrode by using vertically aligned
rutile TiO2 nanowires inlaid with anatase TiO2 nanoparticles for dye-sensitized solar cells. Mater. Chem.
Phys. 143: 1440–1445.
Kontos, A. G., T. Stergiopoulos, G. Tsiminis, Y. S. Raptis, and P. Falaras. 2008. In situ micro- and macro-
Raman investigation of the redox couple behavior in DSSCs. Inorg. Chim. Acta. 361: 761–768.
Koo, H.-J., J. Park, B. Yoo, K. Yoo, K. Kim, and N.-G. Park. 2008. Size-dependent scattering efficiency in
dye-sensitized solar cell. Inorg. Chim. Acta. 361: 677–683.
Krysova, H., J. Trckova-Barakova, J. Prochazka, A. Zukal, J. Maixner, and L. Kavan. 2014. Titania nanofiber
photoanodes for dye-sensitized solar cells. Catal. Today. 230: 234–239.
Kuang, D., C. Klein, H. J. Snaith, R. Humphry-Baker, S. M. Zakeeruddin, and M. Grätzel. 2008. A new ion-
coordinating ruthenium sensitizer for mesoscopic dye-sensitized solar cells. Inorg. Chim. Acta. 361:
699–706.
Kuang, D., C. Klein, Z. Zhang, S. Ito, J.-E. Moser, S. M. Zakeeruddin, and M. Grätzel. 2007. Stable, high-
efficiency ionic-liquid-based mesoscopic dye-sensitized solar cells. Small. 3: 2094–2102.
Kubo, W., T. Kitamura, K. Hanabusa, Y. Wada, and S. Yanagida. 2002. Quasi-solid-state dye-sensitized solar
cells using room temperature molten salts and a low molecular weight gelator. Chem. Commun. 4:
374–375.
Kumara, G. R. A., S. Kaneko, A. Konno, M. Okuya, K. Murakami, B. Onwona-agyeman, and K. Tennakone.
2006. Large area dye-sensitized solar cells: Material aspects of fabrication. Prog. Photovoltaics: Res.
Appl. 14: 643–651.
Kumara, G. R. A., M. Okuya, K. Murakami, S. Kaneko, V. V. Jayaweera, and K. Tennakone. 2004. Dye-
sensitized solid-state solar cells made from magnesium oxide-coated nanocrystalline titanium dioxide
films: Enhancement of the efficiency. J. Photochem. Photobiol. A: Chem. 164: 183–185.
Lagref, J.-J., Md. K. Nazeeruddin, and M. Grätzel. 2008. Artificial photosynthesis based on dye-sensitized
nanocrystalline TiO2 solar cells. Inorg. Chim. Acta. 361: 735–745.
Lai, H., Y. Wang, G. Du, W. Li, and W. Han. 2014. Dual functional YVO4: Eu3+, Bi3+@SiO2 submicron-
sized core–shell particles for dye-sensitized solar cells: Light scattering and downconversion. Ceram.
Int. 40: 6103–6108.
Lee, S., J. Y. Kim, K. S. Hong, H. S. Jung, J.-K. Lee, and H. Shin. 2006. Enhancement of the photoelectric per-
formance of dye-sensitized solar cells by using a CaCO3-coated TiO2 nanoparticle film as an electrode.
Sol. Energy Mater. Solar Cells. 90: 2405–2412.
Lee, W. J., E. Ramasamya, D. Y. Lee, and J. S. Song. 2008b. Performance variation of carbon counter elec-
trode based dye-sensitized solar cell. Sol. Energy Mater. Solar Cells. 92: 814–818.
Lee, W., J. Y. Seng, and J.-I. Hong. 2013. Metal-free organic dyes with benzothiadiazole as an internal accep-
tor for dye-sensitized solar cells. Tetrahedron. 69: 9175–9182.
Lee, Y.-L., Y.-J. Shen, and Y.-M. Yang. 2008a. A hybrid PVDF-HFP/nanoparticle gel electrolyte for dye-
sensitized solar cell applications. Nanotechnology. 19: 455201.
Li, B., G. Lü, L. Luo, and Y. Tang. 2010. TiO2@MgO core-shell film: Fabrication and application to
dye-sensitized solar cells. Wuhan Univ. J. Nat. Sci. 15: 325–329.
Dye-Sensitized Solar Cells 109

Li, P., Y. Duan, Q. Tang, B. He, and R. Li. 2015. An avenue of expanding triiodide reduction and shortening
charge diffusion length in solid-state dye-sensitized solar cells. J. Power Sources. 273: 180–184.
Li, Q., Q. Tang, H. Chen, H. Xu, Y. Qin, B. He, et al. 2014. Quasi-solid-state dye-sensitized solar cells from hydropho-
bic poly(hydroxyethylmethacrylate/glycerin)/polyaniline gel electrolyte. Mater. Chem. Phys. 144: 287–292.
Li, S., Z. Chen, Z. Zhang, Y. Wang, B. Xu, T. Li, and W. Zhang. 2013. One-step electrophoretic deposi-
tion of magnesium oxide-coated tin oxide film and its application to flexible dye sensitized solar cells.
J. Electrochem. Soc. 160: H513–H517.
Li, S.-L., K.-J. Jiang, K.-F. Shao, and L.-M. Yang. 2006. Novel organic dyes for efficient dye-sensitized solar
cells. Chem. Commun. 2792–2794.
Li, Y., J. Hagen, W. Schaffrath, P. Otschik, and D. Haarer. 1999. Titanium dioxide films for photovoltaic cells
derived from a sol-gel process. Sol. Energy Mater. Solar Cells. 56: 167–174.
Li. Z.-Y., M. S. Akhtar, J. H. Kuk, B.-S. Kong, and O.-B. Yang. 2012. Graphene application as a counter elec-
trode material for dye-sensitized solar cell. Mater. Lett. 86: 96–99.
Lin, C.-C., H.-C. Chen, Y. L. Tsai, H.-V. Han, H.-S. Shih, Y.-A. Chang, et al. 2012. Highly efficient CdS-
quantum-dot-sensitized GaAs solar cells. Optics Express. 20: A319–A326.
Lin, L.-Y., M.-H. Yeh, C.-P. Lee, J. Chang, A. Baheti, R. Vittal, et al. 2014. Insights into the co-sensitizer
adsorption kinetics for complementary organic dye-sensitized solar cells. J. Power Sources. 247:
906–914.
Lindström, H., A. Holmberg, E. Magnusson, L. Malmqvist, and A. Hagfeldt. 2001. A new method to make
dye-sensitized nanocrystalline solar cells at room temperature. J. Photochem. Photobiol. A: Chem. 145:
107–112.
Liu, Q.-P. 2014a. Analysis on dye-sensitized solar cells based on Fe-doped TiO2 by intensity-modulated pho-
tocurrent spectroscopy and Mott–Schottky. Chin. Chem. Lett. 25: 953–956.
Liu, Q.-P. 2014b. Photovoltaic performance improvement of dye-sensitized solar cells based on Mg-doped
TiO2 thin films. Electrochim. Acta. 129: 459–462.
Lotey, G. S., and N. K. Verma. 2014. Synthesis and characterization of BiFeO3 nanowires and their applica-
tions in dye-sensitized solar cells. Mater. Sci. Semicond. Proc. 21: 206–211.
Lu, L., R. Li, K. Fan, and T. Peng. 2010. Effects of annealing conditions on the photoelectrochemical proper-
ties of dye-sensitized solar cells made with ZnO nanoparticles. Solar Energy. 84: 844–853.
Luz, A., J. Conradt, M. Wolff, H. Kalt, and C. Feldmann. 2013. p-DSSCs with BiOCl and BiOBr semiconduc-
tor and polybromide electrolyte. Solid State Sci. 19: 172–177.
Ma, T., M. Akiyama, E. Abe, and A. Imai. 2005. High-efficiency dye-sensitized solar cell based on a nitrogen-
doped nanostructured titania electrode. Nano Lett. 5: 2543–2547.
Manfredi, N., A. Bianchi, V. Causin, R. Ruffo, R. Simonutti, and A. Abbotto. 2014. Electrolytes for quasi
solid-state dye-sensitized solar cells based on block copolymers. J. Polym. Sci. A: Poly. Chem. 52:
719–727.
Maruthamuthu, P., S. Fiechter, and H. Tributsch. 2006. Lipid (detergent)-based composite-dye solar cell.
C. R. Chimie. 9: 684–690.
Massihi, N., M. R. Mohammadi, A. M. Bakhshayesh, and M. Abdi-Jalebi. 2013. Controlling electron injection
and electron transport of dye-sensitized solar cells aided by incorporating CNTs into a Cr-doped TiO2
photoanode. Electrochim. Acta. 111: 921–929.
Matsui, H., K. Okada, T. Kawashima, T. Ezure, N. Tanabe, R. Kawano, and M. Watanabe. 2004. Application
of an ionic liquid-based electrolyte to a 100 mm × 100 mm sized dye-sensitized solar cell. J. Photochem.
Photobiol. A: Chem. 164: 129–135.
Matsumura, M., S. Matsudaira, H. Tsubomura, M. Takata, and H. Yanagida, 1980. Dye sensitization and sur-
face structures of semiconductor electrodes. Ind. Eng. Chem. Prod. Res. Dev. 19: 415–421.
Matteocci, F., S. Casaluci, S. Razza, A. Guidobaldi, T. M. Brown, A. Reale, et al. 2014. Solid state dye solar
cell modules. J. Power Sources. 246: 361–364.
Mazille, F., Z. Fei, D. Kuang, D. Zhao, S. M. Zakeeruddin, M. Grätzel, and P. J. Dyson. 2006. Influence of
ionic liquids bearing functional groups in dye-sensitized solar cells. Inorg. Chem. 45: 1585–1590.
Meng, Q.-B., K. Takahashi, X.-T. Zhang, I. Sutanto, T. N. Rao, O. Sato, et al. 2003. Fabrication of an efficient
solid-state dye-sensitized solar cell. Langmuir. 19: 3572–3574.
Michaleviciute, A., M. Degbia, A. Tomkeviciene, B. Schmaltz, E. Gurskyte, J. V. Grazulevicius, J. Bouclé, and
F. Tran-Van. 2014. Star-shaped carbazole derivative based efficient solid-state dye sensitized solar cell.
J. Power Sources. 253: 230–238.
Mihalache, I., A. Radoi, M. Mihaila, C. Munteanu, A. Marin, M. Danila, et al. 2015. Charge and energy
transfer interplay in hybrid sensitized solar cells mediated by graphene quantum dots. Electrochim.
Acta. 153: 306–315.
110 Solar Energy Conversion and Storage

Millington, K. R., K. W. Fincher, and A. L. King. 2007. Mordant dyes as sensitisers in dye-sensitised solar
cells. Sol. Energy Mater. Solar Cells. 91: 1618–1630.
Miyasaka, T., Y. Kijitori, T. N. Murakami, M. Kimura, and S. Uegusa. 2002. Efficient nonsintering type dye-
sensitized photocells based on electrophoretically deposited TiO2 layers. Chem. Lett. 31: 1250–1251.
Mohmeyer, N., D. Kuang, P. Wang, H.-W. Schmidt, S. M. Zakeeruddin, and M. Grätzel. 2006. An efficient
organogelator for ionic liquids to prepare stable quasi-solid-state dye-sensitized solar cells. J. Mater.
Chem. 16: 2978–2983.
Moser, J. 1887. Note on reinforcement photoelectric currents by optical sensibilisirung. Monatsh. Chem. 8: 373.
Murakami, T. N., and M. Grätzel. 2008. Counter electrodes for DSC: Application of functional materials as
catalysts. Inorg. Chim. Acta. 361: 572–580.
Nazeeruddin, M. K., F. D. Angelis, S. Fantacci, A. Selloni, G. Viscardi, P. Liska, et al. 2005. Combined
experimental and DFT-TDDFT computational study of photoelectrochemical cell ruthenium sensitiz-
ers. J. Am. Chem. Soc. 127: 16835–16847.
Nazeeruddin, M., K. P. Péchy, and M. Grätzel. 1997. Efficient panchromatic sensitization of nanocrystalline
TiO2 films by a black dye based on atrithiocyanato-ruthenium complex. Chem. Commun. 1705–1706.
Nishikitani, Y., T. Kubo, and T. Asano. 2006. Modeling of photocurrent in dye-sensitized solar cells fabricated
with PVDF-HFP-based gel-type polymeric solid electrolyte. C. R. Chimie. 9: 631–638.
Noh, Y., and O. Song. 2014. Properties of a Ru/Ti bilayered counter electrode in dye sensitized solar cells.
Electronic Mater. Lett. 10: 271–273.
Olsen, E., G. Hagen, and S. E. Lindquist. 2000. Dissolution of platinum in methoxy propionitrile containing
LiI/I2. Sol. Energy Mater. Solar Cells. 63: 267–273.
O’Regan, B. C., and J. R. Durrant. 2009. Kinetic and energetic paradigms for dye-sensitized solar cells:
Moving from the ideal to the real. Acc. Chem. Res. 42: 1799–1808.
O’Regan, B. C., and M. Grätzel. 1991. A low-cost, high-efficiency solar cell based on dye sensitized colloidal
TiO2 films. Nature. 353: 737–740.
Otaka, H., M. Kira, K. Yano, S. Ito, H. Mitekura, T. Kawata, and F. Matsui. 2004. Multi-colored dye-sensi-
tized solar cells. J. Photochem. Photobiol. A: Chem. 164: 67–73.
Palomares, E., J. N. Clifford, S. A. Haque, T. Lutz, and J. R. Durrant. 2002. Slow charge recombination
in dye-sensitised solar cells (DSSC) using Al2O3 coated nanoporous TiO2 films. Chem. Commun. 14:
1464–1465.
Park, G. W., C. G. Hwang, J. W. Jung, and Y. M. Jung. 2012. Improvement in long-term stability and photovol-
taic performance of UV cured resin polymer gel electrolyte for dye sensitized solar cell. Bull. Korean
Chem. Soc. 33: 4093–4097.
Park, N.-G., J. V.-de Lagemaat, and A. J. Frank. 2000. Comparison of dye-sensitized rutile- and anatase-based
TiO2 solar cells. J. Phys. Chem. B. 104: 8989–8994.
Park, S. K., T. K. Yuna, J. Y. Baea, and Y. S. Wonb. 2013. Combined embedding of N-doping and CaCO3
surface modification in the TiO2 photoelectrodes for dye-sensitized solar cells. App. Surf. Sci. 285:
789–794.
Passoni, L., F. Ghods, P. Docampo, A. Abrusci, J. Martí-Rujas, M. Ghidelli, et al. 2013. Hyperbranched quasi-
1D nanostructures for solid-state dye-sensitized solar cells. ACS Nano. 7: 10023–10031.
Paulsson, H., A. Hagfeldt, and L. Kloo. 2003. Molten and solid trialkylsulfonium iodides and their polyiodides
as electrolytes in dye-sensitized nanocrystalline solar cells. J. Phys. Chem. B 107: 13665–13670.
Pazoki, M., P. W. Lohse, N. Taghavinia, A. Hagfeldt, and G. Boschloo. 2014. The effect of dye coverage on
the performance of dye-sensitized solar cells with a cobalt-based electrolyte. Phys. Chem. Chem. Phys.
16: 8503–8508.
Pichot, F., J. R. Pitts, and B. A. Gregg. 2000. Low-temperature sintering of TiO2 colloids: Application to flex-
ible dye-sensitized solar cells. Langmuir. 16: 5626–5630.
Rai, P., R. Khan, K.-J. Ko, J.-H. Lee, and Y.-T. Yu. 2014. CeO2 quantum dot functionalized ZnO nanorods
photoanode for DSSC applications. J. Mater. Sci.: Mater. Electronics. 25: 2872–2877.
Ramasamy, K., B. Tien, P. S. Archana, and A. Gupta. 2014. Copper antimony sulfide (CuSbS2) mesocrystals:
A potential counter electrode material for dye-sensitized solar cells. Mater. Lett. 124: 227–230.
Ranjusha, R. P., K. R. V. Lekha, Subramanian, V. Nair Shantikumar, and A. Balakrishnan. 2011. Photoanode
activity of ZnO nanotube based dye-sensitized solar cells. J. Mater. Sci. Technol. 27: 961–966.
Rho, W.-Y., J.-W. Choi, H.-Y. Lee, S. Kyeong, S. H. Lee, H. S. Jung, et al. 2014. Dye-sensitized solar cells with silica-
coated quantum dot-embedded nanoparticles used as a light-harvesting layer. New J. Chem. 38: 910–913.
Roh, D. K., W. S. Chi, H. Jeon, S. J. Kim, and J. H. Kim. 2014. High efficiency solid-state dye-sensitized
solar cells assembled with hierarchical anatase pine tree-like TiO2 nanotubes. Adv. Funct. Mater. 24:
379–386.
Dye-Sensitized Solar Cells 111

Rühle, S., M. Shalom, and A. Zaban. 2010. Quantum-dot-sensitized solar cells. Chem. Phys. Chem. 11:
2290–2304.
Saito, Y., W. Kubo, T. Kitamura, Y. Wada, and S. Yanagida. 2004. I−/I3− redox reaction behavior on poly(3,4-
ethylenedioxythiophene) counter electrode in dye-sensitized solar cells. J. Photochem. Photobiol. A:
Chem. 164: 153–157.
Sakaguchi, S., H. Ueki, T. Kato, T. Kado, R. Shiratuchi, W. Takashima, K. Kaneto, and S. Hayase. 2004. Quasi-
solid dye sensitized solar cells solidified with chemically cross-linked gelators: Control of TiO2/gel elec-
trolytes and counter Pt/gel electrolytes interfaces. J. Photochem. Photobiol. A: Chem. 164: 117–122.
Sapp, S. A., C. M. Elliott, C. Contado, S. Caramori, and C. A. Bignozzi. 2002. Substituted polypyridine
complexes of cobalt(II/III) as efficient electron-transfer mediators in dye-sensitized solar cells. J. Am.
Chem. Soc. 124: 11215–11222.
Sarkar, S., A. Makhal, K. Lakshman, T. Bora, J. Dutta, and S. Kumar Pal. 2012. Dual-sensitization via elec-
tron and energy harvesting in CdTe quantum dots decorated ZnO nanorod-based dye-sensitized solar
cells. J. Phys. Chem. C. 116: 14248–14256.
Sayama, K., H. Sugihara, and A. Hironori. 1998. Photoelectrochemical properties of a porous Nb2O5 electrode
sensitized by a ruthenium dye. Chem. Mater. 10: 3825–3832.
Sayama, K., S. Tsukagoshi, K. Hara, Y. Ohga, A. Shinpou, Y. Abe, S. Suga, and H. Arakawa. 2002.
Photoelectrochemical properties of J aggregates of benzothiazole merocyanine dyes on a nanostruc-
tured TiO2 film. J. Phys. Chem. B. 106: 1363–1371.
Schmidt-Mende, L., U. Bach, R. Humphry-Baker, T. Horiuchi, H. Miura, S. Ito, S. Uchida, and M. Grätzel.
2005. Organic dye for highly efficient solid-state dye-sensitized solar cells. Adv. Mater. 17: 813–815.
Senevirathna, M. K. I., P. K. D. D. P. Pitigala, E. V. A. Premalal, K. Tennakone, G. R. A. Kumara, and A.
Konno. 2007. Stability of the SnO2/MgO dye-sensitized photoelectrochemical solar cell. Sol. Energy
Mater. Solar Cells. 91: 544–547.
Shalan, A. E., and M. M. Rashad. 2013. Incorporation of Mn2+ and Co2+ to TiO2 nanoparticles and the perfor-
mance of dye-sensitized solar cells. Appl. Surf. Sci. 283: 975–981.
Shaterian, M., M. Barati, K. Ozaee, and M. Enhessari. 2014. Application of MnTiO3 nanoparticles as coating
layer of high performance TiO2/MnTiO3 dye-sensitized solar cell. J. Ind. Eng. Chem. 20: 3646–3648.
Shi, J., J. Chen, Y. Li, Y. Zhu, G. Xu, and J. Xu. 2015. Three-dimensional network electrolytes with highly
efficient ion-transporting channels for quasi-solid-state dye-sensitized solar cells. J. Power Sources.
282: 51–57.
Shi, J., S. Peng, J. Pei, Y. Liang, F. Cheng, and J. Chen. 2009. Quasi-solid-state dye-sensitized solar cells
with polymer gel electrolyte and triphenylamine-based organic dyes. ACS Appl. Mater. Interfaces. 1:
944–950.
Shinpuku, Y., F. Inui, M. Nakai, and Y. Nakabayashi. 2011. Synthesis and characterization of novel
cyclometalatediridium(III) complexes for nanocrystalline TiO2-based dye-sensitized solar cells.
J. Photochem. Photobiol. A: Chem. 222: 203–209.
Smith, W. 1873. Effect of light on selenium during the passage of an electric current. Nature. 7: 303.
Sun, M., P. Song, J. Li, and X. Cui. 2013. Preparation, characterization and applications of novel carbon and
nitrogen codoped TiO2 nanoparticles from annealing TiN under CO atmosphere. Mater. Res. Bull. 48:
4271–4276.
Suzuki, K., M. Yamaguchi, S. Hotta, N. Tanabe, and S. Yanagida. 2004. A new alkyl-imidazole polymer
prepared as an inonic polymer electrolyte by in situ polymerization of dye sensitized solar cells.
J. Photochem. Photobiol. A: Chem. 164: 81–85.
Suzuki, K., M. Yamaguchi, M. Kumagai, N. Tanabe, and S. Yanagida. 2006. Dye-sensitized solar cells with
ionic gel electrolytes prepared from imidazolium salts and agarose. C. R. Chim. 9: 611–616.
Suzuki, K., M. Yamaguchi, M. Kumagai, and S. Yanagida. 2003. Application of carbon nanotubes to counter
electrodes of dye-sensitized solar cells. Chem. Lett. 32: 28–29.
Tamilavan, V., A.-Y. Kim, H.-B. Kim, M. Kang, and M. H. Hyun. 2014. Structural optimization of thiophene-
(N-aryl)pyrrole-thiophene-based metal-free organic sensitizer for the enhanced dye-sensitized solar cell
performance. Tetrahedron. 70: 371–379.
Tennakone, K., P. K. M. Bandaranayake, P. V. V. Jayaweera, A. Konno, and G. R. R. A. Kumara. 2002. Dye-
sensitized composite semiconductor nanostructures. Physica E. 14: 190–196.
Tennakone, K., A. R. Kumarasinghe, G. R. R. A. Kumara, K. G. U. Wijayantha, and P. M. Sirimanne. 1997.
Nanoporous TiO2 photoanode sensitized with the flower pigment cyanidin. J. Photochem. Photobiol.
A: Chem. 108: 193–195.
Thomas, S., T. G. Deepak, G. S. Anjusree, T. A. Arun, S. V. Nair, and A. S. Nair. 2014. A review on counter
electrode materials in dye-sensitized solar cells. J. Mater. Chem. A 2: 4474–4490.
112 Solar Energy Conversion and Storage

Tian, J., Q. Zhang, L. Zhang, R. Gao, L. Shen, S. Zhang, X. Qu, and G. Cao. 2013. ZnO/TiO2 nanocable
structured photoelectrodes for CdS/CdSe quantum dot co-sensitized solar cells. Nanoscale. 5: 936–943.
Tian, H., L. Hu, C. Zhang, W. Liu, Y. Huang, L. Mo, L. Guo, J. Sheng, and S. Dai. 2010. Retarded charge
recombination in dye-sensitized nitrogen-doped TiO2 solar cells. J. Phys. Chem. C. 114: 1627–1632.
Törngren, B., K. Akitsu, A. Ylinen, S. Sandén, H. Jiang, J. Ruokolainen, et al. 2014. Investigation of plas-
monic gold–silica core–shell nanoparticle stability in dye-sensitized solar cell applications. J. Colloid
Interface Sci. 427: 54–61.
Usui, H., H. Matsui, N. Tanabe, and S. Yanagida. 2004. Improved dye-sensitized solar cells using ionic nano-
composite gel electrolytes. J. Photochem. Photobiol. A: Chem. 164: 97–101.
Wang, C., F. Meng, T. Wang, T. Ma, and J. Qiu. 2014d. Monolithic coal-based carbon counter electrodes for
highly efficient dye-sensitized solar cells. Carbon. 67: 465–474.
Wang, G., Z. Cai, F. Li, S. Tan, S. Xie, and J. Li. 2014a. 2% ZnO increases the conversion efficiency of TiO2
based dye sensitized solar cells by 12%. J. Alloy. Compd. 583: 414–418.
Wang, H., W. Wei, and Y. H. Hu. 2013a. Efficient ZnO-based counter electrodes for dye-sensitized solar cells.
J. Mater. Chem. A. 1: 6622–6628.
Wang, H.-T., D. K. Mishra, P. Chen, and J.-M. Ting. 2014b. p-Type dye-sensitized solar cell based on nickel
oxide photocathode with or without Li doping. J. Alloy. Compd. 584: 142–147.
Wang, P., S. M. Zakeeruddin, P. Comte, I. Exnar, and M. Grätzel. 2003b. Gelation of ionic liquid-based elec-
trolytes with silica nanoparticles for quasi-solid-state dye-sensitized solar cells. J. Am. Chem. Soc. 125:
1166–1167.
Wang, P., S. M. Zakeeruddin, I. Exnar, and M. Grätzel. 2002. High efficiency dye-sensitized nanocrystalline
solar cells based on ionic liquid polymer gel electrolyte. Chem. Commun. 2972–2973.
Wang, P., S. M. Zakeeruddin, J.-E. Moser, and M. Grätzel. 2003a. A new ionic liquid electrolyte enhances the
conversion efficiency of dye-sensitized solar cells. J. Phys. Chem. B. 107: 13280–13285.
Wang, P., S. M. Zakeeruddin, J.-E. Moser, R. Humphry-Baker, and M. Grätzel. 2004b. A solvent-free, SeCN−/
(SeCN)3− based ionic liquid electrolyte for high-efficiency dye-sensitized nanocrystalline solar cells.
J. Am. Chem. Soc. 126: 7164–7165.
Wang, Q., W. M. Campbell, E. E. Bonfantani, K. W. Jolley, D. L. Officer, P. J. Walsh, et al. 2005b. Efficient
light harvesting by using green Zn-porphyrin-sensitized nanocrystalline TiO2 films. J. Phys. Chem. B.
109: 15397–15409.
Wang, S. H., X. W. Zhou, X. R. Xiao, Y. Y. Fang, and Y. Lin. 2014c. An increase in conversion efficiency of
dye-sensitized solar cells using bamboo-type TiO2 nanotube arrays. Electrochim. Acta. 116: 26–30.
Wang, Y. F., J. M. Zhang, X. R. Cui, P. C. Yang, and J. H. Zeng. 2013b. A novel organic ionic plastic crystal
electrolyte for solid-state dye-sensitized solar cells. Electrochim. Acta 112: 247–251.
Wang, Z. S., H. Kawauchi, T. Kashima, and H. Arakawa. 2004a. Significant influence of TiO2 photoelectrode
morphology on the energy conversion efficiency of N719 dye-sensitized solar cell. Coord. Chem. Rev.
248: 1381–1389.
Wang, Z.-S., K. Sayama, and H. Sugihara. 2005c. Efficient eosin Y dye-sensitized solar cell containing Br−/
Br3− electrolyte. J. Phys. Chem. B. 109: 22449–22455.
Wang, Z.-S., T. Yamaguchi, H. Sugihara, and H. Arakawa. 2005a. Significant efficiency improvement of the
black dye-sensitized solar cell through protonation of TiO2 films. Langmuir. 21: 4272–4276.
Wolfbauer, G., A. M. Bond, J. C. Eklund, and D. R. MacFarlane. 2001. A channel flow cell system specifically
designed to test the efficiency of redox shuttles in dye sensitized solar cells. Sol. Energy Mater. Solar
Cells. 70: 85–101.
Wongcharee, K., V. Meeyoo, and S. Chavadej. 2007. Dye-sensitized solar cell using natural dyes extracted
from rosella and blue pea flowers. Sol. Energy Mater. Solar Cells. 91: 566–571.
Wu, M., and T. Ma. 2014. Recent progress of counter electrode catalysts in dye-sensitized solar cells. J. Phys.
Chem. C. 118: 16727–16742.
Wu, S., H. Han, Q. Tai, J. Zhang, S. Xu, C. Zhou, et al. 2008. Enhancement in dye-sensitized solar cells based
on MgO-coated TiO2 electrodes by reactive DC magnetron sputtering. Nanotechnology. 19: 215704.
Wu, W.-Q., Y.-F. Xu, H.-S. Rao, H.-L. Feng, C.-Y. Su, and D.-B. Kuang. 2014. Constructing 3D branched
nanowire coated macroporous metal oxide electrodes with homogeneous or heterogeneous composi-
tions for efficient solar cells. Angew. Chem. Int. Ed. 53: 4816–4821.
Xu, J., M. Li, L.Wu, Y. Sun, L. Zhu, S. Gu, et al. 2014. A flexible polypyrrole-coated fabric counter electrode
for dye-sensitized solar cells. J. Power Sources. 257: 230–236.
Xu, P., W. Meng, F. X. Qing, Z. C. Neng, H. Z. Peng, and D. S. Yuan. 2013. Ionic liquid crystal-based elec-
trolyte with enhanced charge transport for dye-sensitized solar cells. Sci. China. Chem. 56: 1463–1469.
Dye-Sensitized Solar Cells 113

Xue, Z., C. Jiang, L. Wang, W. Liu, and B. Liu. 2014. Fabrication of flexible plastic solid-state dye-sensitized
solar cells using low temperature techniques. J. Phys. Chem. C. 118: 16352–16357.
Yang, J., C. Bao, K. Zhu, T. Yu, F. Li, J. Liu, et al. 2014. High catalytic activity and stability of nickel sulfide
and cobalt sulfide hierarchical nanospheres on the counter electrodes for dye-sensitized solar cells.
Chem. Commun. 50: 4824–4826.
Yodyingyong, S., Q. Zhang, K. Park, C. S. Dandeneau, X. Zhou, D. Triampo, et al. 2010. ZnO nanoparticles
and nanowire array hybrid photoanodes for dye-sensitized solar cells. Appl. Phys. Lett. 96: 73115.
Yue, G.-T., J.-H. Wu, Y.-M. Xiao, J.-M. Lin, M.-L. Huang, L.-Q. Fan, et al. 2013. A dye-sensitized solar cell
based on PEDOT: PSS counter electrode. Chin. Sci. Bull. 58: 559–566.
Yum, J.-H., P. Chen, M. Grätzel, and M.-K. Nazeeruddin. 2008. Recent developments in solid-state dye-
sensitized solar cells. ChemSusChem. 1: 699–707.
Yum, J.-H., S.-S. Kim, D.-Y. Kim, and Y.-E. Sung. 2005. Electrophoretically deposited TiO2 photo-electrodes
for use in flexible dye-sensitized solar cells. J. Photochem. Photobiol. A: Chem. 173: 1–6.
Yum, J.-H., S. Nakade, D.-Y. Kim, and S. Yanagida. 2006. Improved performance in dye-sensitized solar cells
employing TiO2 photoelectrodes coated with metal hydroxides. J. Phys. Chem. B. 110: 3215–3219.
Yusoff, A., N. T. R. N. Kumara, A. Lim, P. Ekanayake, and K. U. Tennakoon, 2014. Impacts of temperature
on the stability of tropical plants pigments as sensitizer for dye sensitized solar cells. J. Biophys. 2014:
739514.
Zhang, G., H. Bala, Y. Cheng, D. Shi, X. Lv, Q. Yu, and P. Wang. 2009. High efficiency and stable dye-sensi-
tized solar cells with an organic chromophore featuring a binary π-conjugated spacer. Chem. Commun.
16: 2198–2200.
Zhang, D., S. M. Lanier, J. A. Downing, J. L. Avent, J. Lum, and J. L. McHale. 2008. Betalain pigments for
dye-sensitized solar cells. J. Photochem. Photobiol. A: Chem. 195: 72–80.
Zhang, D. W., X. D. Lia, H. B. Lia, S. Chena, Z. Suna, X. J. Yinb, et al. 2011. Graphene-based counter elec-
trode for dye-sensitized solar cells. Carbon. 49: 5382–5388.
Zhang, J., S. Najmaei, H. Lin, and J. Lou. 2014. MoS2 atomic layers with artificial active edge sites as transpar-
ent counter electrodes for improved performance of dye-sensitized solar cells. Nanoscale. 6: 5279–5283.
Zheng, Y.-Z., H. Ding, X. Tao, and J.-F. Chen. 2014. Investigation of iodine dopant amount effects on dye-
sensitized hierarchically structured ZnO solar cells. Mater. Res. Bull. 55: 182–189.
Zhou, H., L. Wu, Y. Gao, and T. Ma. 2011. Dye-sensitized solar cells using 20 natural dyes as sensitizers.
J. Photochem. Photobiol. A: Chem. 219: 188–194.
7 Photogalvanic Cells
Yasmin, Abhilasha Jain, Pinki B. Punjabi, and Suresh C. Ameta

CONTENTS
7.1 Introduction........................................................................................................................... 115
7.1.1 Cell Efficiency........................................................................................................... 117
7.1.2 Requirements for the Efficient Photogalvanic Cell................................................... 117
7.2 Dyes....................................................................................................................................... 119
7.3 Complexes.............................................................................................................................. 124
7.4 Miscellaneous........................................................................................................................ 129
References....................................................................................................................................... 132

7.1 INTRODUCTION
The world is experiencing adverse consequences from using current commercial energy resources
that are based on fossil and nuclear fuels. Attention needs to focus on research to find clean and
renewable energy sources. One suitable source of energy in this context is photogalvanic cells.
They are also called liquid-junction solar cells. They can generate electrical energy from solar
energy as well as electrochemical energy and provide the basis for a system with an energy storage
component.
Photogalvanic cells offer a promising area for exploration of the direct use of sunlight. Rideal
and Williams (1925) first discovered the photogalvanic effect, but it was systematically investigated
by Rabinowitch (1940a, 1940b) for an iron–thionine system. Weber and Matijević (1947) made an
attempt to systematize the phenomena in terms of the Becquerel effect. Becquerel’s photogalvanic
effect has been thoroughly investigated on systems composed of different organic redox-dyes and
organic acceptors, especially the speed of changes in potential and the influence of reducing and
oxidizing agents.
Some photogalvanic cells using the iron–thionine system as the photosensitive fluid were built
and tested to explore this suggestion. The observed maximum power conversion efficiency was
3 × 10 –4%, depending on light absorbed. The principal reason for this low efficiency may be polar-
ization of the polished platinum electrodes. Coating these electrodes with platinum black reduced
polarization sufficiently; as a result, it was possible to achieve an efficiency of 6 × 10 –2%, although
this value was not actually observed. It may be possible to further increase efficiency by increasing
electrode area and decreasing electrolyte resistance (Potter and Thaller 1959).
A photogalvanic device or cell is defined as a battery, where the cell solution absorbs light
directly to generate photochemical species which, upon back-reaction through an external circuit
in the presence of suitable electrodes, produce electrical power. It may store a significant amount
of energy as chemical potential under open-circuit conditions and release it as electricity, when the
external circuit is closed. This cell functions as a light recharged storage battery, if the chemical
transformations occur in it without significant degradation over a number of cycles. Such photogal-
vanic cells based upon light-sensitive materials in solution are distinguished from photovoltaic cells,
which are based on inorganic semiconductors (purely solid-state electronic devices). The photovol-
taic devices depend on direct excitation of electrons and their separation from their geminate holes
to produce electrical currents. They have a demerit that they lack inherent capacity for storage. All
practical systems in use for direct conversion of sunlight into electricity utilize solid-state photovol-
taic devices, particularly silicon p-n junctions.

115
116 Solar Energy Conversion and Storage

In the
Illuminated
dark

hν e– e–

hν Cathode
Anode

FIGURE 7.1  Photogalvanic cell.

In the conventional photogalvanic cell (Figure 7.1), a set of electrodes is immersed in an electro-


lyte solution, which contains some light-sensitive materials such as organic dye, complex, and so
on. When the solution in the vicinity of one electrode in the photogalvanic cell is irradiated with
light at a suitable wavelength, and the solution in the vicinity of the second electrode is kept in dark,
then the different energy states of electrochemical equilibrium in the light and dark portions of the
solution will develop a potential difference between these two electrodes. As a result, a shift in elec-
trochemical equilibrium at the electrodes will generate an electrical current in the cell. Thus, the
photogalvanic cell based on light-sensitive materials in electrolyte solution can be used indefinitely,
because its function depends only on reversible shifts in electrical equilibrium in the cell solution in
light and dark. In this cell, no material is consumed as fuel.
The following reactions take place in a photogalvanic cell at electrode and in solution:
In solution,


A+ Z ↽
!!!!
!!!
Dark
! B +Y (7.1)

At electrode,

B↽
!!! A+e
!⇀

(7.2)

Y + e− ↽
!!! Z (7.3)
!⇀

Almost all photogalvanic cells are based upon these photochemical electron-transfer reactions,
because these are relatively free from any irreversible side reactions and yield electrode-responsive
products. The photoredox process involved is

⎯→ B +Y (ΔG > 0) (7.4)


A + Z ⎯hν

The process may be initiated by absorption of light by A, Z, or a sensitizer. Then A/B and Z/Y are
respective redox couples. The A is reduced to B by accepting an electron, while Z is oxidized. The
reverse processes will take place in a reversible shift in the electrochemical equilibrium:

A + e− ↽
!!! B (7.5)
!⇀

Y + e− ↽
!!! Z (7.6)
!⇀

The standard redox potential (Eo) of redox couple A/B must be less than that of the other redox
couple Y/Z. The permissible electrode reactions are represented by Equations 7.5 and 7.6, although
Photogalvanic Cells 117

only certain combinations of these can produce electrical power. The cell reaction should ideally be
the reverse of Equation 7.4:

→ A + Z(ΔG > 0) (7.7)


B +Y ⎯K⎯

It should occur only at the electrode. This is a spontaneous process and occurs before B and Y can
diffuse through solution and react at the electrodes. This homogeneous back-reaction may greatly
reduce the efficiency of the cell. The photogalvanic cell will have good storage capacity, if the back-
reaction is slow enough so that B and Y could be separated or it is so slow that B and Y could coexist
in appreciable concentration in solution. In other words, light energy could be converted to chemical
energy on a recharging process, and chemical energy to electrical energy on discharging process.

7.1.1  Cell Efficiency


It has been predicted theoretically that the power conversion efficiency of a photogalvanic cell could
be appreciably as large as 18%, but the experimental conversion efficiencies are much lower than
expected. The major reasons for this are back-electron transfer, low stability of dyes, and the aggre-
gation of dye molecules around the electrode. It is very difficult to fulfill all the necessary conditions
for better efficiency. It has been estimated that the maximum power conversion efficiency achieved
from a photogalvanic cell lies somewhere between 5% and 9%.
Real systems will exhibit certain extrinsic, nonideal, and device-dependent losses, and as a
result, such systems will exhibit comparatively lower efficiencies.

7.1.2  Requirements for the Efficient Photogalvanic Cell


The performance of photogalvanic cells for the direct conversion of solar energy to electrical energy
depends upon several factors, for example, solution photochemistry, other homogeneous kinetics,
mass transport, electrode kinetics, and load on the cell.
The electrode behavior should be selective, and the photogalvanic cells may be classified in two
categories:

• Concentration cell, where the electrodes are identical, and both electrodes are reversible
to one couple (A/B) and irreversible to the other (Y/Z). Here, the concentration ratio of B
to A is different in two compartments (i.e., dark and illuminated chambers). Therefore, the
cell yields power on irradiation. However, the maximum power point is quite close to the
short-circuit condition, so it is not possible to draw significant currents from this type of
concentration cell, and at the same time, it maintains very different concentrations at the
two electrodes (Figure 7.2).
• Differential cell, where the illuminated electrode is reversible to A/B and irreversible to
Y/Z, but the dark electrode is either reversible to Y/Z and irreversible to A/B or reversible to
both couples. This type of cell gives much better performance in principle. If all the light is
absorbed close to the illuminated electrode, then these two possibilities produce identical
performances (Figure 7.3).

To be useful for conversion and storage of solar energy, there are some requirements for selecting a
photochemical reaction:

• The reaction must be endergonic (energy storing).


• The reaction must be reversible (or cyclic). The backward reaction should be slow enough
to make long-term storage possible under ambient conditions, but it should proceed under
specific conditions to release the energy stored as and when needed.
118 Solar Energy Conversion and Storage

e–

Translucent
electrode Dark electrode
(Reversible to A, B; (Reversible to A, B;
irreversible to Y, Z) irreversible to Y, Z)
Dark solution
Illuminated
(A, Z, Y)
solution
(B, Z, Y)


A+Z B+Y
Dark

A + e–

A + e–

FIGURE 7.2  Concentration cell.

e–

Translucent
electrode Dark electrode
(Reversible to A/B; (Reversible to A/B; irreversible
irreversible to Y/Z) or reversible to Y/Z)
Dark solution
Illuminated
(A, Z, Y)
solution
(B, Z, Y)


A+Z B+Y
Dark

A + e–

Y + e–

FIGURE 7.3  Differential cell.

• The side reactions must be absent, which may lead to the photochemical degradation (irre-
versible) of the reactants.
• Such photochemical reactions that proceed smoothly at any reasonable temperature higher
than the ambient temperature should be investigated.
• It is not necessary for the system to absorb the light radiation directly. Under such circum-
stances, a sensitizer may be used with the absorption bands in the visible region and more
preferably in the near-infrared region.
• The photoproducts of the reaction must be easily separable and should be convenient to
store and/or transport.
• The reactants should be as low cost as possible, even though these are used again and again
in a cyclic process with a forward photoreaction and a backward thermal reaction.
Photogalvanic Cells 119

• The inorganic substrates are preferred over organic compounds, because the organic reac-
tion is normally associated with some side reactions. The organic compounds are con-
sumed during the progress of the reaction; thus, coordination compounds will have an edge
over organic compounds in the future.
• The liquid and solid reactants and products are preferred over gaseous reactants because of
size considerations; however, the gaseous products are not excluded.
• The reactants must be nontoxic, and the reactions should remain unaffected by the atmo-
spheric oxygen.

Photosynthesis is the only photochemical storage system available in nature that can satisfy almost all
of these conditions; despite efforts made by various scientists all over the world to mimic this process
under laboratory conditions, this has not been accomplished. However, some, but not all, photovoltaic
systems satisfy most of these requirements for direct conversion of solar energy into electricity.
A number of compounds have been investigated, and the search is still on, but no system could
fulfill all of the desired conditions. The system fulfilling a majority of these conditions will be pre-
ferred over other systems for photochemical conversion and storage of solar energy. Different sen-
sitizers, including dyes and complexes, have been used in photogalvanic cells by various workers.

7.2 DYES
Clark and Eckert (1975) reported a photogalvanic cell based on the iron–thionine system, where the
power conversion efficiency for absorbed monochromatic light is 1.5%. Gomer (1975) carried out
an analysis of photogalvanic cells and applied it to a specific system, iron–thionine, with only slight
simplifications. A general treatment has been developed for calculating the change in potential at
the point of zero current, when a photogalvanic cell containing two electrochemical redox couples
(such as the iron–thionine system) was illuminated. Potential change on illumination from a ther-
modynamic dark potential to a mixed potential may be positive, negative, or zero. It depends on the
balance between the electrode kinetics of the two couples in the solution. The rate of transport to
the electrode and homogeneous back-reaction in solution also affects the magnitude of the potential
shift (Albery and Archer 1976).
Sensitization of a totally illuminated thin-layer photogalvanic cell based on the iron-thiazine
photoredox reaction has been demonstrated by both current action spectra and enhanced white
light output. Photogalvanic output has been obtained upon illumination of a single solution con-
taining two photoredox dyes and three sensitizers with monochromatic light in the wavelength
range of ~375–700 nm. Sensitization with rhodamine 6 G considerably enhanced cell output under
illumination with white light. However, white light output was inhibited with some sensitizers
(Wildes et al. 1977).
Photoredox reactions and photogalvanic effects of gel systems containing thionine, Fe(II)
salt, and gelatinizing agents were studied by using a thin-layer photocell, with SnO2 and Pt elec-
trodes. The photopotential and photocurrent of such a gel system depends on the characteristics of
the gelatinizing agent. Their values were relatively large and did not decrease with illumination
time, when the gelatinizing agent had intramacromolecular hydrogen bonds. The reason for these
large values and good electron recycling inside the cell may be due to the hydrogen transfer–type
electron-exchange reaction between Fe(III) and Fe(II). It was concluded that the primary structure
of the gelatinizing agent was the most important factor determining the frequency of the electron-
exchange reaction (Shigehara et al. 1978).
Tien et al. (1978) proposed and tested a novel photoelectrochemical cell based on a combined
principle of photogalvanic and photovoltaic effects. The principal element of this cell consists of
a pigmented membrane separating two aqueous solutions, one of which contains thionine dye and
ferrous ions. It was observed that the photo-electromotive force (emf) generated across the cell was
equal to the sum of the voltages derived from photogalvanic and photovoltaic processes. Hall et al.
120 Solar Energy Conversion and Storage

(1977) reported a thin-layer photogalvanic cell based on the iron–thionine system with an efficiency
of 0.03%. The functioning of this cell depends on the selectivity of the tin oxide anode, which
responded preferentially to the thionine–leucothionine couple. The selective response of this elec-
trode was explained in terms of electronic energy levels of semiconductor bands and redox species.
Albery et al. (1979a) used thionine-coated electrode and observed that up to 20 monolayers of
thionine can be reversibly coated on platinum or tin oxide electrode. The kinetics of this thionine–
leucothionine couple was reversible or quasi-reversible. It was shown that reduction of Fe(III) is
much slower on a coated electrode, which means that thionine-coated electrode is one of the ideal
electrodes for a cell.
Tsubomura et al. (1979) studied the photovoltages and photocurrents in various photogalvanic
cells containing a dye (thionine, riboflavin, or proflavin) and a reducing agent (Fe2+, hydroquinone,
EDTA, or triethanolamine). The results indicate that those systems having EDTA or triethanol-
amine as a reducing agent gave much higher photocurrents than those having Fe2+ or hydroquinone.
Such a difference may be due to irreversible reactions of EDTA or triethanolamine taking place
after their photooxidation.
The theory of the operation of the ideal photogalvanic cell for solar energy conversion has been
described, and the crucial kinetic characteristics that the system must possess were deduced for
the homogeneous kinetics, the mass transfer, and the electrode kinetics (Albery and Foulds, 1979).
Solar and Getoff (1979) investigated photophysical and chemical processes affecting the stability
of the thiazine dye–iron system as a photogalvanic cell. Photoelectrochemical devices can be used
under suitable conditions for the production of hydrogen.
Albery et al. (1979b) used the transparent disc electrode to investigate photogalvanic cells using
three possible systems: the iron–thionine system and two iron–ruthenium systems. The modulated
photocurrent was produced when the light source was modulated. Brokken-Zijp and De Groot
(1980) analyzed the kinetics of four reactions of ferrous-thionine photogalvanic cell, which take
place during the illumination. It was concluded that the pseudo-exponential description that was
given in the literature is incorrect. Archer et al. (1980) reported the electrochemical behavior of the
iron–thionine system in acid aqueous sulfate solution in dark. This study was carried out on the
bare platinum electrode and the multilayer thionine-coated platinum. The ferric/ferrous couple has
shown quasi-reversible electrode kinetics at the platinum covered with a submonolayer of thionine,
which was reversibly adsorbed. It was found to be quite irreversible at a thionine-coated electrode.
Albery et al. (1980) synthesized and investigated six new modified thiazine dyes to meet the
photochemical and kinetic requirements of the photogalvanic cell, because efficiency of the iron-
thionine photogalvanic cell for solar energy conversion was severely limited due to the insolubility
of thionine. They explained that for efficient energy conversion, it is also necessary to have a selec-
tive electrode and show that for the thionine-coated electrode it was suitable for the cell.
Fox and Singletary (1980) made use of chemically modified electrodes in dye-sensitized pho-
togalvanic cells. They observed the photocurrents when arene-derivatized electrodes were used as
anodes in an aqueous rhodamine B–hydroquinone photogalvanic cell. These are consistent with
electron injection into the semiconductor from either an excited state or a reduced photoproduct.
The attached molecules presumably function only as energy or electron relays.
Albery and Foulds (1981) studied the suitability of new methylene blue N,N as a dye for the pho-
togalvanic cell. Low quantum efficiency of this type of cell was observed, which may be due to the
dimerization in addition to the diffusion-controlled self-quenching by monomers and dimers. They
concluded that this dye is unsuitable, and that self-quenching may be a serious problem for other
Fe-thiazine systems. The open-circuit voltage of the ferrous-thionine photogalvanic cell has been
reported (Brokken-Zijp et al. 1981). Here, Pt or Au electrode was used in the illuminated chamber.
It was found that experimental open-circuit voltage versus light intensity curves were in agreement
with the calculated values.
Albery et al. (1981) gave two reasons for the poor performance of the iron–thionine photogalvanic
cell: (1) the low solubility of thionine, which means that the incident solar radiation was absorbed
Photogalvanic Cells 121

too far from the illuminated electrode; and (2) the formation of dimers, which do not undergo the
desired photoredox reaction. They reported synthesis and photoelectrochemical characterization
of several isomeric disulfonated thionines. All the isomers examined were sufficiently soluble to
absorb solar radiation close to the illuminated electrode in a photogalvanic cell.
Tamilarasan and Natarajan (1981) observed that thionine dye oxidizes Fe2+ ions in solution on
excitation by light, and the photoproducts undergo dark reactions to restore the starting materials.
The photogalvanic (photovoltaic) potentials were generated by this cyclic process by a cell with
two platinum electrodes, keeping one illuminated and the other in the dark chamber. Light was
converted more effectively into electricity in the totally illuminated thin-layer cell with platinum
and tin oxide electrodes, but the efficiency of the conversion was limited by energy-wasting back-
reactions in solution and by a restriction on the concentration of the dye. They showed that these
limitations can be overcome if thick films (~10 μm) of thionine condensed with macromolecules are
coated onto inert electrodes.
Sharon et al. (1981) studied a thionine/Fe2+/Fe3+ system and found it to be nonreversible after
30 min of exposure in solar radiation. A detailed study was made on a Fe2+/Rh-B system. They
reported that the cell is fully reversible in solar photons giving 40–35 mV potential. This cell can be
charged in 29 min. The rate of discharge in the dark was slow. The mechanism of photoreaction of
the role of rhodamine-B dye was discussed.
Murthy et al. (1982) determined the photovoltages and photocurrent in a photogalvanic cell con-
taining flavin mononucleotide and a reducing agent, ethylenediaminetetraacetic acid. The efficiency
of the cell has been estimated as 0.048%. Srivastava et al. (1982) studied the photogalvanic behavior
of iron–thionine and iron–methylene blue systems in the presence of surfactants. Their data indicate
that the surfactant micelles can be utilized for the storage of light energy. Murthy and Reddy (1983b)
also studied the electrochemical behavior of the systems containing toluidine blue by cyclic voltam-
metry. The photovoltage and photocurrent of this cell containing dye toluidine blue and different
reducing agents like EDTA, triethanol amine, triethylamine, and ferrous ions were measured, and it
was found that photo-outputs with EDTA and amines were higher than with Fe(II).
Roy and Aditya (1983) set up a photoelectrochemical cell based on the photochemistry of anthra-
quinone-2-sulfonate (D). They reported that in the presence of formate at pH 11.0, D on illumination
produces D−• or D2−. At the platinum electrode, the anodic reaction is D−• → D + e−, or D2− → D + 2e−
in the absence of oxygen. At the dark electrode, the cathodic reaction is O2 + 2H2O + 4e− → 4OH−.
The open-circuit potential of the cell is 500 mV, and the short-circuit current is 180 μA. This cell
has been recycled at least eight times. The efficiency of the cell increases with platinized platinum
electrode in the dark chamber. The steady current under illumination is 65 μA with the same open-
circuit voltage of 500 mV. The short-circuit current is 250 μA. The efficiency was even better with a
CdS electrode in the illuminated chamber. Here, the open-circuit voltage is 560 mV. After charging
the cell by illumination for 8 h, a steady current of 120 μA can be drawn from the cell, with illumina-
tion off, for 40 h. The short-circuit current of this cell was 450 μA. The maximum power output was
found to be 4.2 × 10 −6 W. The cell can be recycled at least four times without any loss in efficiency.
Gray deposition on the CdS electrode indicates possible electrode decomposition.
The effect of temperature on phenosafranine-EDTA photogalvanic cell has been studied by
Rohatgi-Mukherjee et al. (1983). They showed that the photovoltage and photocurrent increase lin-
early with increasing temperature, and the time to attain both the equilibrium values also diminished
gradually. Photovoltage and photocurrent appear only in unstirred solutions. The power conversion
efficiency was only 0.24 μW cm–2, and the solar energy efficiency was 10 –3%. The current–potential
curve indicates high activation overpotential, which is responsible for the low efficiency of the cell.
Pan et al. (1983) developed a photoelectrochemical cell involving thiazine photoreduction by
EDTA as the electron donor together with photosynthetic electron transport in chloroplasts to gen-
erate electricity with a higher power conversion efficiency than that obtained for thiazine-EDTA
alone. They reported that the thiazine photogalvanic cell can be combined with the chloroplast solar
battery in order to increase the power conversion efficiency of the system.
122 Solar Energy Conversion and Storage

The improved photo-output in the Fe-thionine photogalvanic cell with polypyrrole-coated selec-
tive electrodes was reported by Murthy and Reddy (1983a). It was found that these electrodes are
sensitive to the ferrous/ferric couple only. Various dyes like thionine, oxonine, Nile blue A, and neu-
tral red modified electrodes have been developed for the photogalvanic cell (Bauldreay and Archer
1983). It was found that the coating of these dyes forms more easily in acidic media than in aprotic
solvents. Acridine yellow/EDTA/K2PtCl6 has proved to be an excellent system for the photoproduc-
tion of hydrogen (Bi and Tien 1984). Further enhancement can be achieved with the addition of
Triton X-100 in this system. Naman and Karim (1984) studied photogalvanic cells using identical
Pt electrodes for different dyes (thionine, methylene blue, eosin, lumiflavine, rhodamine 6G, and
rhodamine B). They recorded the maximum sunlight engineering efficiency (SEE) for these cells
and studied the possibility of a multidye photogalvanic cell.
A quantitative study of electrode selectivity in the performance of a Fe(II)/thionine cell has been
carried out by Groenen et al. (1985) by using various carbon materials as illuminated electrodes.
The open-circuit voltage of this photogalvanic cell depends strongly on the type of carbon electrode
used. These electrodes generated high open-circuit voltage and were selective due to the suppres-
sion of the ferric reduction. The photogalvanic effect of phenosafranine (PSF) dye in an aqueous
solution containing three types of surfactants—CTAB (cationic), SLS (anionic), and Triton X-100
(neutral)—has been investigated (Rohatgi-Mukherjee et al. 1985). It was found that only the PSF/
Triton X-100 system showed a photogalvanic effect.
Aliwi et al. (1985) studied the photogalvanic effect in 13 different organic dyes and vanadium(III)
tris(acetylacetonate) in acetonitrile aqueous solution in a photogalvanic cell. They observed the
effect of acetonitrile concentration for these dyes (new methylene blue, thionine, methylene blue,
eosin, safranine T, rhodamine B, eriochrome black T, Bengal rose, eriochrome blue black B, erio-
chrome red B, vitamin B2, proflavine, and eriochrome blue black R). The sunlight engineering
efficiencies (SEEs) were determined for each dye, and it was found that the best SEE was obtained
when new methylene blue was used at concentration 1 × 10 −5 M. The V(III) complex concentration
was 1 × 10 −4 M at pH 4, and the acetonitrile solution was 40%. They evaluate the reversibility of
this system by monitoring the open-circuit current as a function of time.
Riefkohl et al. (1986) reported results on the dark electrochemistry and photoelectrochemistry of
thionine and two of its disulfonated derivatives (2,6-DST and 4,6-DST) on the electrode (RODRE),
which clearly indicate that coating the electrode with a thin layer of polymeric thionine is crucial to
measuring sizable photoinduced ring currents. The parent thionine does not require the polymeric coat
because, due to its adsorption and aggregation properties, a one- to two-monolayer film of aggregated
molecules spontaneously coats the electrode. They reported that the measured efficiency of the parent
iron–thionine system drops by a factor of 19 or more when semitransparent gold electrodes coated
with thick polythionine layers were utilized in experimental photogalvanic cells. A phenosafranine-
amine photoelectrochemical cell has been studied by Bhowmik and Roy (1987). The photovoltage
generated in these systems was found to correlate with the ionization potential of the amine used.
The thionine-EDTA system has been used for solar energy conversion by Ameta et al. (1989a),
where a photocurrent of 80 µA and photovoltage of 900 mV were obtained. The effects of various
working parameters on the electric output of this cell have also been studied. The photogalvanic
effect of the viologen in the presence of oxygen has been studied by Kaneko and Wohrle (1991).
This viologen cation radical is formed electrochemically on the surface of indium tin oxide–coated
electrode dipped in aqueous electrolyte. The mechanism for generation of photocurrent in the reac-
tion of the photoexcited cation radical with oxygen was proposed.
A novel photoelectrochemical cell has been developed by Bhattacharya et al. (1991), which con-
sists of dye photosafranine and EDTA in various surfactants solution, like CTAB (cationic), SLS
(anionic), and Triton X-100 (neutral); this solution was separated from a saturated aqueous solution
of iodine by Pyrex-sintered glass membrane.
Different redox couples have been used to construct a photogalvanic cell by Jana et al. (1993).
In the construction of these types of cells, semiconductor (InO3)/phenosafranine dye and EDTA
Photogalvanic Cells 123

aqueous solution were placed in one compartment of an H-shaped cell, and the other compartment
contained Cu+/Cu2+, I−/I2, Fe(CN)64−/ Fe(CN)63−, or Fe2+/Fe3+ redox couples. These cells showed two
to three times greater efficiency than the same cell with illuminated Pt electrode. A higher photo-
voltage was reported for a 10-layer cell arrangement of type SnO2/10 mM phenosafranine + 10 mM
brilliant cresyl blue + 0.5 EDTA/Pt. In a photogalvanic cell with tin oxide (working electrode) and
SCE electrode, the photopotentials of aqueous solutions containing xanthenes dyes and electron
donors were determined by Zhou et al. (1994). The hydrogen production was reported for the sys-
tem containing an aqueous solution of erythrosine and triethanolamine. An aqueous solution of
safranine T and triethanolamine was used in this cell, and it was found that this cell shows three to
five times greater values of Voc and 10 times greater values of Isc compared to without safranine T.
The photoelectrochemical behavior of inclusion complexes of dye with β-cyclodextrin has been
investigated by Raj and Ramaraj (1996). It was reported that the addition of β-cyclodextrin to the
dye solution resulted in a twofold increase in the photogalvanic output for the dye-Fe2+ photore-
dox system. This can be attributed to the deaggregation of the photoinactive dimer dye into a pho-
toactive monomer dye. The photoelectrochemical properties of thionine dye covalently attached to
poly(acryloamidoglycolic acid) [P-(AGA)] were studied by Vishwanathan and Natarajan (1996). Here,
the electron transfer processes are more reversible for [P-(AGA)]-TH+ in homogeneous solution.
The photogalvanic effect of Fe(II)-β-diketonate/thionine systems in aqueous acetonitrile was
investigated by Hamdi and Aliwi (1996). They studied the photogalvanic effect of ferrous bis-
(acetylacetonate) (Fe(II)(acac)2) and ferrous bis-(trifluoroacetylacetonate) (Fe(II)(tfac)2) complexes
in aqueous acetonitrile thionine dye solutions in a photogalvanic cell. They determined the theoreti-
cal sunlight engineering efficiency for both complexes and found that the best SEE was obtained
when Fe(II)(tfac)2 was used at a concentration of 1.5 × 10 –4 mol/dm3 with a thionine concentration
of 1 × 10 –4 mol/dm3 at pH = 4 in 40% aqueous acetonitrile.
Cherepy et al. (1997) have constructed a photoelectrochemical cell by utilizing flavanoid antho-
cyanin dyes along with TiO2 powder, which showed 0.56% efficiency under full sun. Jana and
Bhowmik (1997) studied electrode kinetics of photoinduced redox reactions of phenazine dye-
EDTA aqueous systems at different pH levels. They concluded that the photovoltage growth and
decay curves follow the functional forms related to the relaxation times. The inverse of the relax-
ation time or the rate of electrode reaction was pH dependent under the experimental conditions.
Ghosh and Bhattacharya (2002) carried out spectral studies on the interaction of safranine T,
phloxin, and fluorescein with inorganic ions and observed a photogalvanic effect. They reported that
the system consisting of anionic dye fluorescein with inorganic ions in aqueous solution generates pho-
tovoltage in a photogalvanic cell. Fluorescein dye produces much greater photovoltage compared to
other anionic dyes like phloxin (2,4,5,7-tetrabromo-3′,4′,5′,6′-tetrachlorofluorescein) and cationic dye
safranine T (3,6-diamino-2,7-dimethyl-5-phenylphenazinium chloride). The open-circuit photovoltage,
the short-circuit current, and the solar energy efficiencies of these systems were also determined.
The polarity of the microenvironments within a nafion (Nf) film was studied by electrochemi-
cal and photoelectrochemical techniques using a phenothiazine dye and thionine (TH) as a probe
by Abraham and Ramaraj (2004). They observed that a cathodic photocurrent was observed when
the Nf/TH film was exposed to visible light, whereas an anodic photocurrent was observed for
the TBA-Nf/TH film (TBA, tetrabutylammonium). The observed difference in the polarity of the
photocurrent reiterates the polar and less polar environments experienced by TH in Nf and TBA-Nf
films, respectively. The polarity of the photocurrent changes from anodic to cathodic when the
TBA-Nf/TH film is soaked in the photogalvanic cell solution for longer times.
Jana and Rajavenii (2004) reported that the phenozine dyes such as phenosafranine, safranine-O,
and safranine-T form a 1:1 charge-transfer or electron donor–acceptor complex with Triton X-100. They
observed that the photogalvanic and photoconductivity studies also support the above molecular inter-
action. Some anionic xanthene dyes show enhancement of fluorescence intensity with a red shift and
develop photovoltage in a photoelectrochemical cell in the presence of Triton X-100 (Bhowmik and
Ganguly 2005).
124 Solar Energy Conversion and Storage

Ameta et al. (2006) studied the use of a bromophenol red-EDTA system for generation of electricity
in a photogalvanic cell. The photopotential and photocurrent generated by this cell were 581.0 mV and
45 μA, respectively, while the use of mixed dyes toluidine blue and azur-B system and EDTA as a reduc-
tant in the photogalvanic cell for solar energy conversion and storage was investigated by Lal and Yadav
(2007). The photocurrent and photopotential generated by this system were 40.0  μA and 802.0  mV,
respectively. The observed conversion efficiency was 0.0708%, and maximum output of the cell was
32.08 µW. The cell can be used for 2 h in the dark. Genwa et al. (2009) studied the use of photosensitizers
toluidine blue and malachite green in the presence of NaLS for photogalvanic solar energy conversion.
They found that the conversion efficiency and storage capacity of the developed cells were 0.1448% and
123 min with the toluidine blue system, and were 0.059% and 32 min with the malachite green system.
Halls et al. (2012) studied wastewater as a photoelectrochemical fuel source. They developed a
photogalvanic cell that employed 2,4-dicholorophenol as a fuel source, an N-substituted phenothi-
azine as the light harvester, and a sacrificial zinc anode. This cell has approximately 0.4% light to
electrical power conversion efficiency in violet light. Mahmuoud et al. (2014) observed an improve-
ment in a photogalvanic cell for solar energy conversion and storage using a rose Bengal-oxalic
acid-tween 80 system and studied its commercial viability. The cell, as developed, can work for
230 min in the dark on irradiation for 72 min.
Azmat and Uddin (2012) observed photocurrent during the reduction of methylene blue with
maltose into leucomethylene blue. The dye molecule presumably functions as an energy or elec-
tron relay. Absorption maxima, quantum yields, and lifetimes of the triplet transient species were
reported. Results showed that transmitted light intensity depends upon the concentration of reduc-
ing sugar, solution acidity, and temperature.
Solid-state photogalvanic dye-sensitized solar cells were developed by Berhe et al. (2014). They
examined thermal electron transfer from molecular sensitizers to nanostructured semiconductor
electrodes composed of TiO2 nanorods. They also studied electron-accepting molecular dyes along
with an aryl amine as the electron donor. It was observed that the dyes operating by thermal injec-
tion into TiO2 function work better in solid-state photoelectrochemical cells than in liquid-junction
cells due to the kinetic advantage of solid-state cells with respect to photoinduced acceptor quench-
ing to form the necessary radical anion sensitizers.
Genwa and Chouhan (2004) investigated the use of three heterocyclic dyes, azur A, azur B,
or azur C, in photogalvanic cells for solar energy conversion and storage with a NaLS-ascorbic
acid system. They reported the conversion efficiencies for azur A/azur B/azur C-NaLS-ascorbic
acid systems to be 0.5461%, 0.9646%, and 0.4567% and storage capacities 110, 135, and 95 min,
respectively.
On the basis of experimental observations, a general mechanism for a photogalvanic cell has
been derived. The nature of the electroactive species was confirmed by the effect of diffusion length
on imax and ieq. The reduced form of dye (semireduced or leuco form of dye) has been proposed as
an electroactive species at the illuminated electrode (platinum electrode), while dye itself acted as
an electroactive species in a dark chamber at the counterelectrode (SCE).
Many researchers have tried photogalvanic cells containing different combinations of dyes,
reductants, and/or surfactants. Their results have been summarized in Table 7.1.

7.3 COMPLEXES
Hoffman and Lichtin (1979) examined the photochemical determinants of the efficiency of pho-
togalvanic cell operation. They are (1) the absorption spectral characteristics of the cell solution,
(2) the efficiency of the formation of separated charge carriers, and (3) the lifetimes of the carriers
toward back-electron transfer. They achieved modulation of bulk solution dynamics by variation
of the solution medium and discussed the photochemical determinants with particular reference to
the use of thionine or [Ru(bpy)3]2+ as the light-absorbing species. Current–time and potential–time
characteristics were measured, and the transient processes in the photogalvanic cell containing
Photogalvanic Cells 125

TABLE 7.1
Some Photogalvanic Cells
Photopotential (mV) Conversion Efficiency (%)
System: Photocurrent (µA) Fill Factor Performance
Dye-Reductant-Surfactant Power (µW) in Dark (min) Reference
Methylene 654 — Ameta et al. (1989b)
blue-EDTA-NaLS 190 —
— —
Azur A-glucose-NaLS 811 — Khamesra et al. (1990)
1470 —
— —
Rhodamine 6G-oxalic 414 0.55 Genwa and Mahaveer (2007)
acid-CTAB 90 0.45
25.22 —
Methyl 625 0.2707 Genwa and Khatri (2009)
orange-DTPA-brij-35 95 0.40
28.16 94
Safranine 785 0.9769 Gangotri and Gangotri (2009)
O-EDTA-tween-80 300 0.34
101.6 60
Rhodamine B-DTPA-NaLS 843 0.74 Genwa and Kumar (2010)
185 0.4955
77.28 85
Bromo cresol green- 834 0.80 Genwa and Singh (2013)
ascorbic acid-NaLS 350 0.23
83.52 140
Rhodamine 1162 1.26 Meena et al. (2013)
6G-EDTA-NaLS 510 —
131.6 168
Biebrich scarlet-ascorbic 919 0.8967 Genwa and Sagar (2013)
acid-tween-60 210 —
93.15 75
Rose bengal-oxalic 666 0.258 Mahmoud et al. (2014)
acid-tween-80 119.5 0.328
22.30 230
Azur C-EDTA 879 — Lodha et al. (1991)
75 —

Azur A-NTA 362 — Dube et al. (1993)
60 —
17.82 —
Azur B-NTA 340 — Dube et al. (1997)
60 —
18.75 —
Methylene blue-oxalic acid 312 0.211 Gangotri and Meena (2001)
110 0.28
12.6 35
Safranine T-I– 503 0.0031 Ghosh and Bhattacharya (2002)
2.0 —
— —
(Continued)
126 Solar Energy Conversion and Storage

TABLE 7.1 (Continued)


Some Photogalvanic Cells
Photopotential (mV) Conversion Efficiency (%)
System: Photocurrent (µA) Fill Factor Performance
Dye-Reductant-Surfactant Power (µW) in Dark (min) Reference
Phloxin-Br – 410 0.0027
2.0 —
— —
Fluorescein-Br– 1274 0.0059
4 —
— —
Azur B-NTA 996 0.1685
70 0.25
17.52 12
Rose Bengal-EDTA 520 0.9706 Madhwani et al. (2007a)
56 0.36
29.12 28
Fuchsine basic-EDTA 650 — Madhwani et al. (2007b)
60 —
— —
Fluorescein-EDTA 418 — Madhwani et al. (2007c)
42 —
— —
Brilliant cresyl blue-fructose — — Sharma et al. (2011)
590 —
183.3 —
Eosin-fructose 848 0.84 Gangotri and Bhimwal (2011)
240 0.3404
87.52 55
Fast Green FCF-fructose 1083 1.33 Koli (2014)
431 —
138.6 70
Azur-(mannitol + NTA) 347 — Dube (1993)
80 —
— —
(Toluidine blue + 695 0.16 Lal and Gangotri (2011)
thionine)-EDTA 105 —
72.9 42
(Brilliant cresyl blue + 871 1.87 Gangotri and Mahawar (2012)
toluidine)-ethylene 630 —
glycol-NaLS 548.73 70
(Brilliant green + celestine 636 0.31 Yadav and Lal (2013)
blue)-EDTA 93 —
59.1 65

Ru(bipy)32+ and iron have been studied by Daul et al. (1980). A maximum was observed in the volt-
age-time curve when high light intensity and low to medium work load of the cell was employed.
A photogalvanic cell based on photolysis of rubidium anion in THF has been investigated by
Goldstein et al. (1980). This photogalvanic cell was made up of two half-cells of type [Rb-THF]I
and [Rb- THF, crown ether (or kryptand)]II. Values of 30 nA photocurrent and 250 mV photovoltage
were obtained by irradiation with the wavelength above 320 nm, whereas with the radiations with
wavelength above 620 nm, only 6 nA photocurrent and 80 mV photovoltage were generated.
Photogalvanic Cells 127

A model system for singlet state–driven photogalvanic cells has been reported involving photo-
reduction of 3,7-diaminophenoxazinylium chloride (oxonine) by iron(II) (Creed et al. 1981). Yamase
and Ikawa (1980) observed the pH dependence of the alkylammonium molybdate system and pro-
duction of hydrogen by a photogalvanic cell.
The alternating current photogalvanic cell has been developed by Daul et al. (1981). This cell
has a Ru(bipy)32+/Fe3+ system and symmetrical electrode configuration. Here, the two electrodes
were illuminated by light beams, alternatively, and resulted in almost rectangular current–time
and voltage–time responses. This cell is stable and eliminates the disadvantages of slow diffusion
processes, because mass transfer is not necessary.
Kawai and Yamamura (1982) have developed a thin-layer photocell, where the working electrode
was made up of tin oxide modified with tris(2,2-bipyridine)ruthenium complex immobilized in a
polyvinylcinnamate membrane. Platinum was used as the counterelectrode, and an aqueous solu-
tion containing the disodium salt of EDTA was used as an electrolyte solution. It was also reported
that the addition of the tetracyanoquinodimethane salt (TCNQ) of poly-4-vinylpyridine and neutral
tetracyanoquinodimethane enhanced the generation of photocurrent.
4-Phenylpyridine (as ancillary ligand) in Ru(II) polypyridyl complexes has been used by Garcia
et al. (1998) for sensitization of the n-type TiO2 electrode. Two novel molecular sensitizers, cis-
{(dcbH2)2Ru(ppy)2}2+ and cis-{(dcbH2)2Ru(ppy)H2O}2+ were prepared, where dcbH2 = 4,4′-(CO2H)2-
2,2′-bipyridine and ppy = 4-phenylpyridine. They achieved a higher incident photon-to-current
conversion efficiency value and broader spectral response in longer wavelengths using a derivative
with one coordinated ppy, cis-{(dcbH2)2Ru(ppy)(H2O2+)}. It was observed that the number of azine
ligands coordinated to the nonattached side extends spectral sensitivity to the visible light.
The generation of photocurrent with the graphite electrodes coated with polymer films contain-
ing tris(2,2′-bipyridyl)Ru(II) and viologen was observed by Oyama et al. (1982). Hoffman and Sima
(1983) examined a new technique for photochemical conversion of solar energy based on ligand
photodissociation from metal complexes. The concept was illustrated with a photogalvanic cell,
where voltages are generated by photodissociation of CO from carbonylferroheme, and with a cell,
where the illuminated electrode was coated with an iron tetraphenylporphyrin. A singlet state–
driven photogalvanic cell based on the photoreduction of 3,7-diaminophenoxazinylium chloride
(oxonine) by iron(II) has been investigated by Fawcett et al. (1985).
The photoredox couple comprising vanadium(III) bis (2,2′-bipyridyl) chloride-Fe(III) also
showed photogalvanic effect (Aliwi et al. 1986). It was reported that generation of photocurrent and
photovoltage is directly proportional to the incident light intensity. Casado et al. (1990) investigated
the photogalvanic behavior of K3Mn(CN)6 in aqueous cyanide solution. It was observed that irra-
diation of anodic solution containing Mn(II) complex with UV light resulted in increasing current.
The stationary value of the photocurrent lasted for 6.5 h after putting off the light, when the current
reached its maximum value.
Aliwi and Hanna (1987) determined new Z-values for 20 pure solvents and their aqueous mix-
tures at 25%, 50%, and 75% water by volume by using 1-methyl-4-carbomethoxy-pyridinium iodide
at room temperature. They investigated the effect of pure solvents and their water mixtures on the
VCl3− -thionine photogalvanic system. They determined the voltage efficiency of these photogal-
vanic cells for each solution and correlated that with the Z-values. They reported that the highest
voltage efficiency was obtained with acetonitrile and acetone and aqueous mixtures of these two
solvents.
Operating characteristics of a photogalvanic cell using RS rubrum chromatophores have been
studied by Erabi et al. (1987). They investigated operating characteristics of a photogalvanic cell
using a Pt vertical chromatophore, m-PMS, ascorbate vertical SnO//z system. Open-circuit photo-
voltage and short-circuit photocurrent about 60 mV and 200–500 nA, respectively, were reported.
Photogalvanic generation of dihydrogen by water splitting using [MoS4]2− as catalyst has been
reported by Bhattacharyya et al. (1988). They achieved 4% solar energy efficiency for photogenera-
tion of dihydrogen using [MoS4]2− as catalyst, when the anode compartment of a photogalvanic cell
128 Solar Energy Conversion and Storage

was illuminated. The generated photocurrent rises slowly with time and reaches a limiting value. A
secondary dark reaction sets in after 10 h, which produces H2 and photocurrent, even when the light
was switched off. They also suggested possible mechanisms in both of these cases.
The photogalvanic effect has been reported for alkylpyridinium salt in water or in water–
methanol solution by Markov and Novkirishka (1991). This process was found to be reversible.
Here, the value of the electrode potential depends on the type of salt and solvent used. It was
observed that the absence of atmospheric oxygen in the irradiated solution affects the energy as well
as potential of this photogalvanic cell.
Photogalvanic cells based on thionine-ferrous, [Ru(bipy)3]2+/[PtCl6]2–, [Ru(5-Clphen)3]2+/Fe[(5-
Clphen)3]3+ systems were investigated by Titse et al. (1993). Various methods of electrode surface
chemical modification were applied in order to transfer the photoelectrochemical processes in a
solid phase. Gobi and Ramaraj (1993) studied the electron transfer reaction at Nafion and mont-
morillonite clay coated electrode containing [Ru(bipy)3]2+ in the presence of Fe3+ and HClO4.
These electrodes showed different photoelectrochemical properties depending on the coating mate-
rials. A novel photogalvanic cell has been constructed by using Pt/Nafion-[Ru(bipy)3]2+ and Pt/
MM-[Ru(bipy)3]2+ as a photoanode and photocathode, respectively.
Gobi and Ramaraj (1994) also constructed a photogalvanic cell using Nafion and montmorillon-
ite (MM) clay adsorbed [Ru(bpy)3]2+ coated electrodes in the presence of Fe3+ ions. They observed
that the Nafion and MM clay adsorbed [Ru(bpy)3]2+ electrodes exhibited different photoelectro-
chemical behavior depending on the nature of the coating material. Anodic behavior was observed
at a Nafion-[Ru(bpy)3]2+ electrode with reference to an inert electrode in a photogalvanic cell, while
the MM clay-[Ru(bpy)3]2+ electrode showed cathodic polarity. A new photogalvanic cell was con-
structed by coating one electrode with Nafion-[Ru(bpy)3]2+ and the other electrode with MM clay-
[Ru(bpy)3]2+. They concluded that this new photogalvanic cell showed an additive photogalvanic
response on visible light irradiation.
Suresh et al. (1999) reported manganese-molybdenum-diethyldithiocarbamate complex as a
potential system for solar energy conversion. They reported that the manganese-molybdenum-
diethyldithiocarbamate [MnMoO2(Et2dtc)4(H2O)] {Et2dtc  =  diethyldithiocarbamate} complex
exhibited reversible photogalvanic behavior in aqueous dimethylformamide medium in a Honda
cell. The photogalvanic behavior was observed by varying the pH, temperature, and photosensitiz-
ers. UV, visible, and sunlight were used as the light sources. A potential of 345 mV was obtained at
80°C in visible light, and the system was found to be reversible for several cycles. They constructed
a photoelectrochemical cell by coupling a charged nickel electrode with the complex electrode,
which incorporates the experimental compound in acetylene black on a nickel substrate. They also
reported a maximum potential of 1.08 V with 80 μA current, when irradiated with a tungsten lamp.
Suresh Raj et al. (2000) did photoelectrochemical studies on [MnMoO2(NCS)(Ox)3(H2O)2] (Ox =
8-quinolinol) complex in aqueous dimethylformamide medium in a Honda cell. They reported that
this system developed a maximum potential of 335 mV, when exposed to visible light at 30°C, and
it was reversible. When a temperature difference was maintained between the illuminated and dark
half-cells, this system generated 410  mV at 60°C. They also reported a solid-state galvanic cell,
using the complex mixed with tetraethylammonium perchlorate (TEAP), which showed a maxi-
mum voltage of 25  mV. A sandwich galvanic cell, constructed from transparent tin oxide-glass/
complex/platinum system was also developed with a maximum photovoltage of 88 mV, when irradi-
ated with a tungsten halogen lamp.
Matsumoto et al. (2001) constructed a photogalvanic cell using a photosynthetic reaction cen-
ter complex and cytochrome c2 from Rhodospirillum rubrum. This photosynthetic reaction center
complex was immobilized on a p-benzoquinonethiol-modified Au electrode. The potential of the
reaction center–immobilized Au electrode in 0.1 M (M = mol/dm3) Tris-HCl buffer (pH 7.5) was
shifted to negative direction upon illumination, and the value increased with increasing intensity
of incident light. It was observed that the potential shift was enhanced by adding cytochrome c2 to
the solution. A photogalvanic cell was constructed using this electrode, cytochrome c2, and an ITO
Photogalvanic Cells 129

electrode with hydrophilic surface. The cell performance was examined, where open-circuit photo-
voltage and short-circuit photocurrent were about 5 mV and 40 μA, respectively.
The photogalvanic behavior of [Cr2O2S2(l-Pipdtc)2 (H2O2)] was investigated in a Honda cell using
100% DMF and different percentages of water–DMF systems (Pokhrel and Nagaraja 2009). The max-
imum potential of 200 mV with 18 µA current was generated in DMF. The system was found to be
reversible when the irradiated solution was aerated immediately; the solution was found to be irrevers-
ible when it was kept in the dark for a long time (12 h) and then aerated. The Cr(V) is photoreduced to
Cr(IV) with the light irradiation, and the unstable Cr(IV) reverts to Cr(V) by aerial oxidation.
Long-lived photoinduced charge separation in a [Ru(Bpy)3]2+/viologen system at the Nafion mem-
brane-solution interface was observed by Yi et al. (2000). They examined the photoinduced electron
transfer from the excited state of tris(2,2′-bipyridine)ruthenium(II) [Ru(bpy)3]2+ incorporated into
Nafion membranes to propylviologen sulfonate (PVS) in the surrounding solution by photochemi-
cal and photoelectrochemical measurements. Here, N,N′-tetramethylene-2,2′-bipyridinium [DQ]2+
entrapped in the Nafion membranes was used as an electron relay. An electrode was fabricated by
coating [Ru(bpy)3]2+-[DQ]2+-incorporated Nafion film on an ITO glass. The photoinduced voltage
of this electrode was measured with a saturated calomel reference electrode in PVS solution to be
~350 mV, when the light intensity of the order ~60 mW cm–2 was used. This electrode was used as
the illuminated electrode to construct a photogalvanic cell with a platinum electrode as the dark
electrode. Irradiation of this electrode with visible light results in cathodic photocurrent. There is no
net chemical change associated with the functioning of this cell, which converts light to electricity.

7.4 MISCELLANEOUS
The photoelectrolysis of water has been investigated using cells that consist of an illuminated n-TiO2
(rutile) anode, an aqueous electrolyte, and a platinized-Pt cathode (Mavroides et al. 1975). It has
been found that such cells operate either in the photogalvanic mode (no H2 evolved) or in the pho-
toelectrolytic mode (H2 evolved at the cathode by decomposition of water). Maximum values of
80%–85% for the external quantum efficiency for current production in the photogalvanic mode
have been measured at hν  ≈  4  eV with single-crystal and polycrystalline TiO2 anodes. Similar
results were obtained in preliminary experiments with SrTiO3 anodes. The internal quantum effi-
ciencies, corrected for reflection and absorption losses, are close to 100%. This indicated that the
band bending in TiO2 under photogalvanic conditions was sufficient to separate the electron–hole
pairs generated by photon absorption and also that the oxygen overvoltage for charge transfer at
the semiconductor–electrolyte interface was negligible for illuminated anodes.
Fong and Winograd (1976) reported the photogalvanic effects arising from chlorophyll a-quin-
hydrone (Chl:H2Q, 1:1) half-cell reactions using platinum electrodes. A photocurrent was developed
in the presence of light as Chl a aggregates presumably undergo a charge-transfer interaction that
results in the creation of a p-type semiconductor film. Photopotential at the Pt-Chl a electrode was
positive. When the light was turned off, the half-cells regressed toward the preillumination condi-
tions, and a reverse current was observed. The spectral response of the photogalvanic effect has also
been determined. It was concluded that a distribution of Chl a-H2O aggregates contributes to the
observed photocurrent.
A general treatment has been developed for the current–voltage characteristics of photogalvanic
cells containing two identical electrodes and operating by virtue of homogeneous reactions. This
treatment involved the electrode and homogeneous kinetics, the relative magnitudes of the diffusion
and reaction layers, and the position of the photostationary state. It was observed that power cannot
be drawn from this cell if both the couples were either highly reversible or highly irreversible. The
best performance was obtained if one couple is highly reversible, and the other is highly irreversible.
If the homogeneous kinetics are rapid, an important contribution to the current delivered by the cell
from catalytic currents due to the presence of Y and Z close to the electrode was observed (Albery
and Archer 1977).
130 Solar Energy Conversion and Storage

Zeichner et al. (1978) studied photogalvanic effects in the aqueous ferric bromide system. This
system exhibited a considerable photogalvanic effect, with the illuminated half-cell normally posi-
tive in potential with respect to the nonilluminated one. They determined the photogalvanic open-
circuit potential as a function of solution composition and incident light intensity. The results are in
agreement with the kinetic mechanisms for the photolysis, subsequent reactions, and mixed poten-
tial measurements.
Mountz and Tien (1978) constructed a novel type of photoelectrochemical cell based on a
unique combination of a photogalvanic and photovoltaic effect and demonstrated the feasibility
of photo-galvano-voltaic cell for light conversion. Tien et al. (1979) also described a new type of
electrochemical photocell. It was based on the combined principles of the photovoltaic (PV) and
photogalvanic (PG) phenomena. This system had the advantages of both cells (PV and PG) and is,
therefore, called the photogalvanovoltaic (PGV) cell. The key element of the cell responsible for
the PV effect was a porphyrin-coated glassy carbon electrode. Pt or glassy carbon served as the
counterelectrode.
Fox and Kabir-ud-Din (1979) developed a new carbanionic photogalvanic cell. Cyclooctatetraenyl
dianion functioned as an electron source when excited with visible light at the surface of a TiO2
semiconductor electrode in a liquid NH3 photoelectrochemical cell. Since the conduction band
of TiO2 in ammonia is located below the oxidation potential of disodium or dipotassium salts of
cyclooctatetraene at –0.55 eV versus Ag, the efficiency of the photoinduced electron transfer may
depend on high-lying surface states, kinetic retardation of electron exchanges at the ammonia-
semiconductor interface, or other relaxation phenomena.
The theory of transient processes in photogalvanic cells has been developed by Daul et al. (1979).
The systems consisted of a photoactive species B, which can be oxidized (or reduced) to A in its
excited state B* by the one-electron acceptor Z, which itself is reduced (oxidized) to Y. It was based
on an assumption that both redox couples undergo one-electron transfer reactions at the metal elec-
trodes. The electrode kinetics for both redox couples was taken into account. A computer program
was described, where data such as current, potential of the illuminated electrode, potential of the
dark electrode, power, and the concentration profiles of A, B, Z, and Y near the electrodes can be
obtained as functions of time, after illumination has started.
The relationship between open-circuit photovoltage and light intensity in photogalvanic cells has
been studied by Quickenden and Yim (1979). All the expressions gave logarithmic relationships
between photovoltage and light intensity when the concentration perturbations are small. Under
certain other conditions, the logarithmic expressions approximate linear relationships.
Fox and Singletary (1980b) described the synthesis and the chemical and physical properties of
some arene-derivatized tin oxide semiconductor electrodes. They employed several techniques for
this attachment (esterification, silanation, silanation/amidation, and the use of cyanuric chloride
as a linking agent). The benefits of each technique were evaluated. The relationship between the
observed properties of the attached arenes and the potential utility of the derivatized electrodes
as anodes in photogalvanic cells was discussed. Murthy et al. (1980) examined the photogalvanic
cell, making use of the photoreduction of riboflavin with ethylenediaminetetraacetic acid. They
observed an appreciable photoinduced voltage in this cell, but the photoinduced current was low.
The estimated sunlight engineering efficiency has been estimated to be approximately 0.018%.
Haas (1980) described various processes in the photogalvanic cell and their possibilities in solar
energy storage and also limitations of these cells. The photolysis of various hydrocarbon anions
with the visible light leads to those photoreactions, which have applications in photochemical
energy storage reactions and photogalvanic cells (Fox and Singletary 1980).
The photogalvanic effect has been observed by Katz et al. (1981) during the photooxidation of
pheophytin a in dry acetone. It was observed that the photopotential can be increased by the addi-
tion of Mn2+ ions for this cell. Open-circuit photovoltage of 750 mV and short-circuit current of
100 µA cm–2 can be obtained with the photoanode on the base of pheophytin-sensitized oxidation
of Mn2+ ions.
Photogalvanic Cells 131

Stevenson and Erbelding (1981) reported a photogalvanic cell where visible light was converted
into electricity via the photodissociation of iodine in aqueous and nonaqueous solutions. They
explained a mechanism for the aqueous system that probably involves the formation of I2− or I
radical, which is reduced at the transparent illuminated electrode. The efficiency was sensitive to
formal concentration of I2, and wavelength, and somewhat less on electrode material, with the high-
est efficiency observed (0.03%) in a cell containing indium tin oxide electrode at 4047 Å and 9F I2,
3F NaI in acetonitrile.
Dixit and Mackay (1982) investigated the current and voltage responses of the totally illumi-
nated thin-layer photogalvanic cell in microemulsions, micellar solution, and water. A significant
enhancement of power output in anionic microemulsion was reported as compared to that in cat-
ionic microemulsion, micellar solution, and water. They also examined the power conversion effi-
ciency of the photocell as a function of the various parameters, such as dye concentration, light
intensity, pH, and distance between the electrodes. The highest percent solar engineering efficiency
obtained was 0.33 × 10 –3.
Photogalvanic cells using heteropoly electrolytes have been developed by Papaconstantunou and
Ioannidis (1983). These compounds undergo reduction on illumination with UV-vis light, whereas
oxidation of organic species takes place. The potential difference has been developed between the
photoreduced light half-cell and similar dark half-cell. Groenen et al. (1984) analyzed the effect
of the nonionic amphiphile Triton X-100 micelles on the electrical performance of the ferrous/
thionine photogalvanic cell. They reported that the addition of Triton X-100 micelles to the aqueous
acidic cell solution leads to solubilization of the dye thionine in the outer poly(oxyethylene) spheres
of the micelles. As a result, the solubility of thionine increases, and thermal back-reactions were
suppressed, whereas the diffusion of species toward the electrodes slowed. An overall efficiency
increase of fivefold has been reached relative to the photogalvanic cell, free of micelles.
Kamat (1985, 1986) studied photoelectrochemistry in colloidal systems using a photogalvanic
cell based on TiO2 semiconductor colloid. The feasibility of employing a colloidal semiconductor
system in the operation of a photogalvanic cell was demonstrated. A photogalvanic power conver-
sion efficiency of 0.002% was achieved with thionine and colloidal TiO2 upon bandgap excitation.
The use of TiO2 colloid as a carrier for the deposition of a photoactive or an electroactive species on
the electrode surface was discussed.
A riboflavin-diethanolamine system has also shown photogalvanic effects (Prajuntaboribal and
Chaikum 1987). Photogalvanic properties of an Ag/Langmuir-Blodgett film/GaAs structure based
on the fluorocarbon polymer have been studied by Znamensky et al. (1992). It was reported that the
efficiency of such a type of structure with the thickness of 10–16 nm of Langmuir-Blodgett films
was six times higher than simple Ag/GaAs diodes.
Markov et al. (1996) worked on photopotential and photocurrent induced by an aqueous solution
of aliphatic alcohols. They reported the photogalvanic behavior of aliphatic alcohols in aqueous
solution, where UV illumination caused the appearance of photopotential and photocurrent. The
values of the electrode potential and energy of this photogalvanic cell depend on the type of alcohol
used, its concentration, and the temperature and the pH of the irradiated solution. An important
characteristic of the phenomenon observed was its reversibility. After discharge of a photogalvanic
cell by application of an external load in the circuit, a subsequent irradiation again leads to the
appearance of photoinduced electrode potential and photocurrent.
Mbindyo et al. (1997) prepared and evaluated the performance of a small laboratory-scale reac-
tor employing a nanocrystalline titanium dioxide anode and a platinum black cathode for pollutant
decomposition with simultaneous reduction of water to produce hydrogen. This reactor requires
only light as an energy input, and it operates as a photogalvanic cell to produce electricity. They
achieved oxidative photodegradation of some phenols: 4-chlorophenol, 2,4,5-trichlorophenol, and
4,4′-dichlorobiphenyl. It was observed that the solutions of 0.1 mM chlorophenols were decomposed
in 2–3 h, with an average turnover of 5.4 × 1015 molecules cm–2 s–1, whereas complete degrada-
tion of chlorophenols to carbon dioxide, water, and chloride ion was achieved in less than 6 h.
132 Solar Energy Conversion and Storage

4,4′-Dichlorobiphenyl was poorly water soluble; hence, it was adsorbed onto soil and suspended
in a pH 13 anode solution for its decomposition. Hydrogen gas was produced at the cathode at a
rate 1.4 mL h–1 cm–2 using an anode solution of pH 13 and a cathode solution of pH 1. The average
reactor potential under these conditions was 1.53 V, and the power output was 0.36 mW at a current
density of 2 mA cm–2.
Li et al. (2002) fabricated an electrode by transferring the Langmuir-Blodgett film on an ITO
glass and used this electrode as the illuminated electrode in a photogalvanic cell with a platinum
electrode as the dark electrode. They found that irradiation of this electrode with visible light results
in anodic photocurrent, and there is no net chemical change associated with the function of the cell
converting light to electricity.
Properties and applications of semiconductor-nanoparticle-donor nanochemical composites as
energy sources of the future have been analyzed by Tsivadze et al. (2012). An idea of using nano-
film-nanoparticle-electron donor composites as the energy sources to substitute oil and gas has been
developed. Electrons are knocked out from nanoparticles under the effect of visible light, and are
absorbed by the nanofilm, which acts as a reservoir. Finally, these are attached to the donor, and
then the cycle was repeated. They analyzed the electronic structure, properties, and interactions
between all the nanocomposite components, as well as their interaction with the solvent and protec-
tive shell.
Halls and Wadhawan (2012) developed a lightweight, autonomous, and practical concept for an
electrical power source. The quasi-biphasic, entirely new concept electrochemical cell, based an
electron transfer rather than ion transfer, has been shown to act as a photogalvanic device, where
violet light irradiation exhibited maximum light to electrical power conversion efficiency of ~2%. It
has the additional ability to act as an electrically rechargeable electrochemical capacitor of voltage
efficiency 85% and power efficiency ~80%.
Halls and Wadhawan (2013) also considered a mathematical model for a photosynthesis-inspired
regenerative photogalvanic device. The performance of this system as a solar cell (for cells con-
structed from electrochemically reversible redox couples with fast photoinduced electron transfer
reaction) is critically dependent on the concentration of the supersensitizer. Based on the numerical
simulations, it has been suggested that this regenerative system has a solar-to-electrical power con-
version efficiency of 4.5%, an attractive realistic single cell value.
Although the photogalvanic cells can be used in the dark for a relatively longer time (especially
in the presence of surfactant in the form of micelles), the conversion efficiency of these systems is
too low to make it commercially useful. The search is still on for newer photogalvanic cells with
greater efficiency, but there are many more miles to go before such cells become economically via-
ble and useful for such applications. However, such cells can be used for applications in electronic
systems, where current at the microampere limit is required.

REFERENCES
Abraham, J. S., and R. Ramaraj. 2004. Microenvironment effects on the electrochemical and photoelectro-
chemical properties of thionine loaded Nafion films. J. Electroanal. Chem. 561: 199–126.
Albery, W. J., and M. D. Archer. 1976. Photogalvanic cells-I. The potential of zero current. Electrochim. Acta.
21: 1155–1163.
Albery, W. J., and M. D. Archer. 1977. Photogalvanic cells. 2. Current-voltage and power characterstics.
J. Electrochem. Soc. 124: 688–697.
Albery, W. J., P. N. Bartlett, J. P. Davies, A. W. Foulds, A. R. Hillman, and F. S. Bachiller. 1980. New thiazine
dyes for photogalvanic cells. Faraday Discuss. Chem. Soc. 70: 341–357.
Albery, W. J., P. N. Bartlett, A. W. Foulds, F. A. Souto-Bachiller, and R. Whiteside. 1981. Photogalvanic cells.
Part 14. The synthesis and characterization of disulphonated thionines. J. Chem. Soc., Perkin Trans. 2:
794–800.
Albery, W. J., W. R. Bowen, F. S. Fisher, A. W. Foulds, K. J. Hall, A. R. Hillman, R.l G. Egdell, and A. F.
Orchard. 1979a. Photogalvanic cells: Part XI. The thionine-coated electrode. J. Electroanal. Chem.
Interfacial Electrochem. 107: 37–47.
Photogalvanic Cells 133

Albery, W. J., W. R. Bowen, F. S. Fisher, and A. D. Turner. 1979b. Photogalvanic cells. Part IX. Investigations
using the transparent rotating disc electrode. J. Electroanal. Chem. 107: 11–22.
Albery, W. J., and A. W. Foulds. 1979. Photogalvanic cells. J. Photochem. 10: 41–57.
Albery, W. J., and A. W. Foulds. 1981. Photogalvanic cells XIII: New methylene blue NN. J. Photochem. 15: 321–328.
Aliwi, S. M., and E. M. Hanna. 1987. New measured Z-values and their effects on the voltage efficiency of
VCI3− thionine photogalvanic system. Magallat buhut al-taqat al-Samsiyyat. 5: 39–52.
Aliwi, S. M., S. A. Naman, and I. K. Al-Daghstani. 1985. Photogalvanic effect in organic dyes/vanadium (III)
tris (acetylacetonate) in aqueous acetonitrile solution. Magallat buhut al-taqat al-Samsiyyat. 3: 49–61.
Aliwi, S. M., S. A. Naman, and I. K. Al-Daghstani. 1986. Photogalvanic effect in the vanadium(III) bis(2,2′-
bipyridyl) chloride Fe(III) system. Sol. Cells. 18: 85–91.
Ameta, S. C., S. Khamesra, A. K. Chittora, and K. M. Gangotri. 1989b. Use of sodium lauryl sulphate in a
photogalvanic cell for solar energy conversion and storage: Methylene blue-EDTA system. Int. J. Energy
Res. 13: 643–647.
Ameta, S. C., S. Khamesra, S. Lodha, and R. Ameta. 1989a. Use of the thionine-EDTA system in photogal-
vanic cells for solar energy conversion. J. Photochem. Photobiol. A: Chem. 48: 81–86.
Ameta, S. C., P. B. Punjabi, J. Vardia, S. Madhwani, and S. Chaudhary. 2006. Use of bromophenol red–EDTA
system for generation of electricity in a photogalvanic cell. J. Power Sources. 159: 747–751.
Archer, M. D., M. Isabel, C. Ferreira, W. J. Albery, and A. R. Hillman. 1980. Photogalvanic cells: Part XII.
The dark electrochemistry of the iron-thionine system at platinum. J. Electroanal. Chem. Interfacial
Electrochem. 111: 295–308.
Azmat, R., and F. Uddin. 2012. Photogalvanic effect of maltose/methylene blue system in aqueous methanol.
Asian J. Chem. 24: 2833–2838.
Bauldreay, J. M., and M. D. Archer. 1983. Dye-modified electrodes for photogalvanic cells. Electrochim. Acta.
28: 1515–1522.
Berhe, S. A., H. B. Gobeze, S. D. Pokharel, E. Park, and W. J. Youngblood. 2014. Solid state photogalvanic dye
sensitized solar cells. ACS Appl. Mater. Interfaces. 6: 10696–10705.
Bhattacharya, S., A. K. Jana, and B. B. Bhowmik. 1991. Storage solar cell consisting of phenosafranin-EDTA
in surfactant solution. J. Photochem. Photobiol. A: Chem. 56: 81–87.
Bhattacharyya, R. G., D. P. Mandal, and K. K. Rohatgi Mukherjee. 1988. Photogalvanic generation of dihy-
drogen by water splitting using [MoS4]2− as catalyst. Bull. Mater. Sci. 10: 373–379.
Bhowmik B. B., and P. Ganguly. 2005. Photophysics of xanthene dyes in surfactant solution. Spectrochim.
Acta A. 61: 1997–2003.
Bhowmik, B. B., and S. Roy. 1987. A phenosafranin-amine photoelectrochemical cell. Energy. 12: 519–521.
Bi, Z.-C. and H. T. Tien. 1984. Photoproduction of hydrogen by dye-sensitized systems. Int. J. Hydrogen
Energy. 9: 717–722.
Brokken-Zijp, J. C. M., and M. S. De Groot. 1980. The kinetics of chemical reactions in ferrous-thionine
photogalvanic solutions. Chem. Phys. Lett. 76: 1–6.
Brokken-Zijp, J. C. M., M. S. De Groot, and P. A. J. M. Hendriks. 1981. The open-circuit voltage of the
ferrous-thionine photogalvanic cell. Chem. Phys. Lett. 81: 129–135.
Casado, J., J. Peral, J. Balué, and X. Domenech. 1990. Photogalvanic behaviour of K3Mn(CN)6 in CN− aqueous
solutions. Electrochim. Acta. 35: 427–429.
Cherepy, N. J., G. P. Smestad, M. Grätzel, and J. Z. Zhang. 1997. Ultrafast electron injection: Implications
for a photoelectrochemical cell utilizing an anthocyanin dye-sensitized TiO2 nanocrystalline electrode.
J. Phys. Chem. B. 101: 9342–9351.
Clark, W. D. K., and J. A. Eckert. 1975. Photogalvanic cells. Sol. Energy. 17: 147–150.
Creed, D., N. C. Fawcett, and R. L. Thompson. 1981. Photoreduction of 3,7-diaminophenoxazinylium chlo-
ride (‘oxonine’) by iron(II). A model system for singlet-state driven photogalvanic cells. J. Chem. Soc.,
Chem. Commun. 497–499.
Daul, C., O. Haas, A. Lottaz, A. V. Zelewsky, and H.-R. Zumbrunnen. 1980. Transient processes in photo-
galvanic cells: Part II. The Ru (bipy)32+/Fe3+ cell. J. Electroanal. Chem. Interfacial Electrochem. 112:
51–61.
Daul, C., O. Haas, and A. Von Zelewsky. 1979. Transient processes in photogalvanic cells. Part I. Fundamentals.
J. Electroanal. Chem. 107: 49–58.
Daul, C., O. Haas, A. V. Zelewsky, and Hans-Rudolf Zumbrunnen. 1981. Transient processes in photogal-
vanic cells: Part III. The alternating current photogalvanic cell. J. Electroanal. Chem. Interfacial
Electrochem. 125: 307–313.
Dixit, N. S., and R. A. Mackay. 1982. Microemulsions as photogalvanic cell fluids. The surfactant thionine-
iron(II) system. J. Phys. Chem. 86: 4593–4598.
134 Solar Energy Conversion and Storage

Dube, S. 1993. Simultaneous use of two reductants in a photogalvanic cell for solar-energy conversion and
storage. Int. J. Energy Res. 17: 311–314.
Dube, S., A. Lodha, S. L. Sharma, and S. C. Ameta. 1993. Use of an Azur-A-NTA system in a photogalvanic
cell for solar energy conversion. Int. J. Energy Res. 17: 359–363.
Dube, S., S. L. Sharma, and S. C. Ameta. 1997. Photogalvanic effect in azur B-NTA system. Energy Conver.
Manage. 38: 101–106.
Erabi, T., T. Sengoku, K. Nishimura, and M. Wada. 1987. Operating characteristics of a photogalvanic cell
using RS. Rubrum chromatophores. 1987. Chem. Express. 2: 455–458.
Fawcett, N. C., D. Creed, R. L. Thompson, and D. W. Presser. 1985. A singlet state- driven photogalvanic cell
based on the photoreduction of 3,7-diaminophenoxazinylium chloride (‘oxonine’) by iron(II). J. Chem.
Soc., Chem. Commun. 11: 719–720.
Fong, F. K., and N. Winograd. 1976. In vitro solar conversion after the primary light reaction in photosynthe-
sis. Reversible photogalvanic effects of chlorophyll-quinhydrone half-cell reactions. J. Am. Chem. Soc.
98: 2287–2289.
Fox, M. A., and Kabir-ud-Din. 1979. A new carbanionic photogalvanic cell. J. Phys. Chem. 83: 1800–1801.
Fox, M. A., and N. J. Singletary. 1980. Solar energy utilization by carbanion photolysis. Sol. Energy. 25:
225–229.
Gangotri, K. M., and M. K. Bhimwal. 2011. The photochemical conversion of solar energy into electrical
energy: Eosin-fructose system. Environ. Prog. Sust. Energy. 30: 493–499.
Gangotri, K. M., and A. K. Mahawar. 2012. Comparative study on effect of mixed photosensitizer system
for solar energy conversion and storage: Brilliant cresyl blue + toluidine blue–ethylene glycol–NaLS
system. Environ. Prog. Sust. Energy. 31: 474–480.
Gangotri, K. M., and R. C. Meena. 2001. Use of reductant and photosensitizer in photogalvanic cells for solar
energy conversion and storage: Oxalic acid-methylene blue system. J. Photochem. Photobiol. A: Chem.
141: 175–177.
Gangotri, P., and K. M. Gangotri. 2009. Studies of the micellar effect on photogalvanics: Solar energy conver-
sion and storage in EDTA-Safranine O-Tween-80 system. Energy Fuels. 23: 2767–2772.
Garcia, C. G., N. Y. Murakami Iha, R. Argazzi, and C. A. Bignozzi. 1998. 4-Phenylpyridine as ancillary
ligand in ruthenium(II) polypyridyl complexes for sensitization of n-type TiO2 electrodes. J. Photochem.
Photobiol. A: Chem. 115: 239–242.
Genwa, K. R., and A. Chouhan. 2004. Studies of effect of heterocyclic dyes in photogalvanic cells for solar
energy conversion and storage: NaLS-ascorbic acid system. J. Chem. Sci. 116: 339–345.
Genwa, K. R., and N. C. Khatri. 2009. Use of Brij-35-Methyl orange-DTPA system in photogalvanic cell for
solar energy conversion and storage. Indian J. Chem. Technol. 16: 396–400.
Genwa, K. R., and A. Kumar. 2010. Role of Rhodamine B in photovoltage generation using anionic surfactant
in liquid phase photoelectrochemical cell for solar energy conversion and storage. J. Indian Chem. Soc.
87: 933–939.
Genwa, K. R., A. Kumar, and A. Sonel. 2009. Photogalvanic solar energy conversion: Study with photosensi-
tizers toluidine blue and malachite green in presence of NaLS. Appl. Energy. 86: 1431–1436.
Genwa, K. R., and Mahaveer. 2007. Role of surfactant in the studies of solar energy conversion and storage:
CTAB-rhodamine 6G-oxalic acid system. Indian J. Chem. Sec. A. 46: 91–96.
Genwa, K. R., and C. P. Sagar. 2013. Energy efficiency, solar energy conversion and storage in photogalvanic
cell. Energy Conver. Manage. 66: 121–126.
Genwa, K. R., and K. Singh. 2013. Use of bromocresol green-ascorbic acid-NaLS system in photogalvanic cell
for solar energy conversion and storage. J. Indian Chem. Soc. 90: 813–819.
Getoff, N. 1990. Photoelectrochemical and photocatalytic methods of hydrogen production: A short review.
Int. J. Hydrogen Energy. 15: 407–417.
Ghosh, J. K., and S. C. Bhattacharya. 2002. Spectral studies on the interaction of safranine T, phloxin, fluores-
cein with inorganic ions and their photogalvanic effect. J. Indian Chem. Soc. 79: 225–230.
Gobi, K. V., and R. Ramaraj. 1993. Electron transfer reactions at Nafion and clay adsorbed Ru (bpy)2+3 coated
electrodes. J. Mol. Catal. 84: 187–192.
Gobi, K. V., and R. Ramaraj. 1994. Photoinduced electron transfer reactions of [Ru(bpy)3]2+ adsorbed onto
Nafion® and clay coated electrodes in the presence of Fe3+ ions. J. Electroanal. Chem. 368: 75–85.
Goldstein, S., S. Jaenicke, and H. Levanon. 1980. Photogalvanic cell based on the photolysis of rubidium
anions in THF. Chem. Phys. Lett. 71: 490–493.
Gomer, R. 1975. Photogalvanic cells. Electrochim. Acta. 20: 13–20.
Groenen, E. J. J., M. S. de Groot, and R. de Ruiter. 1985. Carbon electrodes in the ferrous/thionine photogal-
vanic cell: A quantitative study of electrode selectivity. Electrochim. Acta. 30: 1199–1204.
Photogalvanic Cells 135

Groenen, E. J. J., M. S. De Groot, R. De Ruiter, and N. De Wit. 1984. Triton X-100 micelles in the ferrous/
thionine photogalvanic cell. J. Phys. Chem. 88: 1449–1454.
Haas, O. 1980. Processes in photogalvanic cells and their possibilities in solar energy storage In J. Silverman
(Ed.), Energy Storage: A Vital Element in Mankind’s Quest for Survival and Progress. Oxford:
Pergamon, pp. 433–435.
Hall, D. E., W. D. K. Clark, J. A. Eckert, N. N. Lichtin, and P. D. Wildes. 1977. Photogalvanic cell with semi-
conductor anode. Am. Ceram. Soc. Bull. 56: 408–411.
Halls, J. E., T. Johnson, A. A. Altalhi, and J. D. Wadhawan. 2012. Wastewater as a photoelectrochemical fuel
source: Light-to-electrical energy conversion with organochloride remediation. Electrochem. Commun.
22: 4–7.
Halls, J. E., and J. D. Wadhawan. 2012. Photogalvanic cells based on lyotropic nanosystems: Towards the use
of liquid nanotechnology for personalized energy sources. Energy Environ. Sci. 5: 6541–6551.
Halls, J. E. and J. D. Wadhawan. 2013. A model for efficient, semiconductor-free solar cells via supersensitized
electron transfer cascades in photogalvanic devices. Phys. Chem. Chem. Phys. 15: 3218–3226.
Hamdi, S. T., and S. M. Aliwi. 1996. The photogalvanic effect of Fe(II)-β-diketonate/thionine systems in
aqueous acetonitrile. Monatsh. Chem. 127: 339–346.
Hoffman, M. Z., and N. N. Lichtin. 1979. Photochemical determinants of the efficiency of photogalvanic con-
version of solar energy. Trans. J. Brit. Ceram. Soc. 153–187.
Hoffman, B. M., and P. D. Sima. 1983. Solar energy conversion through ligand and photodissociation. J. Am.
Chem. Soc. 105: 1776–1778.
Jana, A. K., and B. B. Bhowmik. 1997. Electrode kinetics of photoinduced redox reactions: Phenazine dye-
EDTA aqueous systems at different pH. J. Photochem. Photobiol. A: Chem. 110: 41–46.
Jana, A. K., and S. Rajavenii. 2004. Studies on the molecular interaction of phenazine dyes with Triton X-100.
Spectrochim. Acta A. 60: 2093–2097.
Jana, A. K., S. Roy, and B. B. Bhowmik. 1993. Photoelectrochemical cells consisting of phenosafranin-EDTA
and different redox couples with illuminated semiconductor electrode. Sol. Energy. 51: 313–316.
Kamat, P. V. 1985. Photoelectrochemistry in colloidal systems. Part 2. A photogalvanic cell based on TiO2
semiconductor colloid. J. Chem. Soc., Faraday Trans. 1. 81: 509–518.
Kamat, P. V. 1986. Erratum: Photoelectrochemistry in colloidal systems. Part 2. A photogalvanic cell based
on TiO2 semiconductor colloid. J. Chem. Soc., Faraday Trans. 1. 82: 1031.
Kaneko, M., and D. Wöhrle. 1991. Novel photogalvanic effect of the viologen cation radical formed electro-
chemically in the presence of oxygen. J. Electroanal. Chem. Interfacial Electrochem. 30: 209–215.
Katz, E. Y., Y. N. Kozlov, and B. A. Kiselev. 1981. Photoanode on the base of pheophytin-sensitized reactions.
Energy Convers. Manage. 21: 171–174.
Kawai, W., and S. Yamamura. 1982. A thin layer photocell based on a tris(2,2′-bipyridine)ruthenium com-
plex immobilized in a membrane containing an electronically conductive polymer. J. Memb. Sci. 12:
107–117.
Khamesra, S., R. Ameta, M. Bala, and S. C. Ameta. 1990. Use of micelles in photogalvanic cell for solar
energy conversion and storage. Azur A-glucose system. Int. J. Energy Res. 14: 163–167.
Koli, P. 2014. Solar energy conversion and storage: Fast green FCF-fructose photogalvanic cell. Appl. Energy.
118: 231–237.
Lal, C., and K. M. Gangotri. 2011. Energy conversion and storage potential of photogalvanic cell based on
mixed dyes system: Ethylene diaminetetraacetic acid-toluidine blue-thionine. Environ. Prog. Sust.
Energy. 30: 754–761.
Lal, C., and S. Yadav. 2007. Use of mixed dyes in photogalvanic cell for solar energy conversion and storage:
EDTA-toluidiene blue and azur-B system. Asian J. Chem. 19: 981–987.
Li, J. Y., M. L. Peng, L. Zhang, L. Zhu Wu, B. Wang, and C. Tung. 2002. Long-lived photoinduced charge
separation in carbazole-pyrene-viologen system incorporated in Langmuir-Blodgett films of substituted
diazacrown ethers. J. Photochem. Photobiol.: A Chem. 150: 101–108.
Lodha, S., S. Khamesra, B. Sharma, and S. C. Ameta. 1991. Use of an Azur C-EDTA system in a photogal-
vanic cell for solar energy conversion. Int. J. Energy Res. 15: 431–435.
Madhwani, S., R. Ameta, J. Vardia, P. B. Punjabi, and V. K. Sharma. 2007c. Fluoroscein-EDTA system in
photogalvanic cell for solar energy conversion. Energy Sources, Part A. 29: 721–729.
Madhwani, S., S. Chaudhary, P. B. Punjabi, V. K. Sharma, and S. C. Ameta. 2007a. Use of rose bengal-EDTA
system for the generation of electricity in a photogalvanic cell. J. Indian Chem. Soc. 84: 181–183.
Madhwani, S., J. Vardia, P. B. Punjabi, and V. K. Sharma. 2007b. Use of fuchsine basic: Ethylenediaminetetraacetic
acid system in photogalvanic cell for solar energy conversion. Proc. Inst. Mech. Engg. Part A: J. Power
Energy. 221: 33–39.
136 Solar Energy Conversion and Storage

Mahmoud, S. A., B. S. Mohamed, A. S. El-Tabei, M. A. Hegazy, M. A. Betiha, H. M. Killa, E. K. Heikal, S.


A. K. Halil, M. Dohium, and S. B. Hosney. 2014. Improvement of the photogalvanic cell for solar energy
conversion and storage: Rose Bengal-oxalic acid-Tween 80 system. Energy Procedia. 46: 227–236.
Markov, P., and M. Novkirishka. 1991. On the photogalvanic properties of alkylpyridinium salts in solution.
Electrochim. Acta. 36: 1287–1289.
Markov, P., M. Novkirishka, and K. Aljanapy. 1996. Photopotential and photocurrent induced by an aqueous
solution of aliphatic alcohols. J. Photochem. Photobiol. A: Chem. 96: 161–165.
Matsumoto, K., S. Fujioka, Y. Mii, M. Wada, and T. Erabi. 2001. Construction of a photogalvanic cell using
photosynthetic reaction center complex and cytochrome c2 from Rhodospirillum rubrum. Electrochem.
69: 340–343.
Mavroides, J. G., D. I. Tchernev, J. A. Kafalas, and D. F. Kolesar. 1975. Photoelectrolysis of water in cells with
TiO2 anodes. Mater. Res. Bull. 10: 1023–1030.
Mbindyo, J. K. N., M. F. Ahmadi, and J. F. Rusling. 1997. Pollutant decomposition with simultaneous genera-
tion of hydrogen and electricity in a photogalvanic reactor. J. Electrochem. Soc. 144: 3153–3158.
Meena, A. S., Rishikesh, and R. C. Meena. 2013. Electrochemical studies of anionic and nonionic micelles
with dyes and reductant in photogalvanic cell. Int. J. Innov. Res. Sci. Eng. Technol. 2: 6118–6123.
Mountz, J. M., and H. T. Tien. 1978. The photogalvanovoltaic cell. Sol. Energy. 21: 291–295.
Murthy, A. S. N., R. Bhargava, and K. S. Reddy. 1982. Flavin mononucleotide (FMN)-ethylenediaminetetra
acetic acid (EDTA) photogalvanic cell. Int. J. Energy Res. 6: 389–395.
Murthy, A. S. N., H. C. Dak, and K. S. Reddy. 1980. Photogalvanic effect in riboflavin-ethylenediaminetet-
raacetic acid system. Int. J. Energy Res. 4: 339–343.
Murthy, A. S. N., and K. S. Reddy. 1983a. Polypyrrole coated selective electrodes for iron-thionine photogal-
vanic cell. Electrochim. Acta. 28: 473–476.
Murthy, A. S. N., and K. S. Reddy. 1983b. Studies on photogalvanic effect in systems containing toluidine
blue. Sol. Energy. 30: 39–43.
Naman, S. A., and A. S. R. Karim. 1984. Efficiency of photogalvanic cells with various dyes. J. Sol. Energy
Res. 2: 31–41.
Oyama, N., S. Yamaguchi, M. Kaneko, and A. Yamada. 1982. Photocurrent generation of graphite electrodes
coated with polymer films confining tris(2,2′-bipyridyl) ruthenium (II) and viologen. J. Electroanal.
Chem. Interfacial Electrochem. 139: 215–222.
Pan, R. L., R. Bhardwaj, and E. L. Gross. 1983. Photochemical energy conversion by a thiazine photosyn-
thetic-photoelectrochemical cell. J. Chem. Technol. Biotechnol. Chem. Technol. 33 A: 39–48.
Papaconstantinou, E., and A. Ioannidis. 1983. Photogalvanic cells using heteropoly electrolytes. Inorg. Chim.
Acta. 75: 235–236.
Pokhrel, S., and K. S. Nagaraja. 2009. Photogalvanic behaviour of [Cr2O2S2(1-Pipdtc)2(H2O)2] in aqueous
DMF. Sol. Energy Mater. Solar Cells. 93: 244–248.
Potter, Jr., A. E., and L. H. Thaller. 1959. Efficiency of some iron-thionine photogalvanic cells. Sol. Energy. 3: 1–7.
Prajuntaboribal, K., and N. Chaikum. 1987. Photogalvanic effect in the riboflavin-diethanolamine system. Sol.
Energy. 38: 149–153.
Quickenden, T. I., and G. K. Yim. 1979. The relationship between open circuit photovoltage and light intensity
in photogalvanic cells—An extension of Albery and Archer’s treatment. Electrochim. Acta. 24: 143–146.
Rabinowitch, E. 1940a. The photogalvanic effect I: The photochemical properties of the thionine-iron system.
J. Chem. Phys. 8: 551–559.
Rabinowitch, E. 1940b. The photogalvanic effect II. The photogalvanic properties of the thionine-iron system.
J. Chem. Phys. 8: 560–566.
Raj, C. R., and R. Ramaraj. 1996. Electrochemistry and photoelectrochemistry of phenothiazine dye-β-
cyclodextrin inclusion complexes. J. Electroanal. Chem. 405: 141–147.
Rideal, E. K., and E. G. Williams. 1925. The action of light on the ferrous iodine iodide equilibrium, J. Chem.
Soc. Trans. 127: 258–269.
Riefkohl, J., L. Rodriquez, L. Romero, and F. Souto. 1986. Photoelectrochemistry with the rotating optical
disk ring electrode: II. The thionine-coated electrode is selective to the leuco forms of the parent thio-
nine and the 2,6- and 4,6-disulfonated thionine derivatives. J. Electrochem. Soc. 133: 1828–1834.
Rohatgi-Mukherjee, K. K., R. Chaudhuri, and B. B. Bhowmik. 1985. Molecular interaction of phenosafranin
with surfactants and its photogalvanic effect. J. Colloid Interface Sci. 106: 45–50.
Rohatgi-Mukherjee, K. K., M. Roy, and B. B. Bhowmik. 1983. Photovoltage generation in the solid phenosaf-
ranine-EDTA sandwich cell. Solar Energy. 31: 417–419.
Roy, A., and S. Aditya. 1983. A novel photogalvanic cell using anthraquinone-2-sulfonate. Int. J. Hydrogen
Energy. 8: 91–96.
Photogalvanic Cells 137

Sharma, U., P. Kohli, and K. M. Gangotri. 2011. Brilliant cresyl blue-fructose for enhancement of solar energy
conversion and storage capacity of photogalvanic solar cells. Fuel. 90: 3336–3342.
Sharon, M., S. G. Sharan, A. Sinha, and B. M. Prasad. 1981. Saur vidyut kosh-solar photogalvanic cell-II.
J. Electrochem. Soc. India. 30: 200–203.
Shigehara, K., M. Nishimura, and E. Tsuchida. 1978. Photogalvanic effect of thin-layer photocells composed
of thionine/Fe(II) systems. Electrochim. Acta. 23: 855–860.
Solar, S., and N. Getoff. 1979. Photophysical and chemical processes affecting the stability of the thiazine
dye-iron system. Int. J. Hydrogen Energy. 4: 403–410.
Srivastava, R. C., P. R. Marwadi, P. K. Latha, and S. B. Bhise. 1982. Solar energy storage using surfactant
micelles. Int. J. Energy Res. 6: 247–251.
Stevenson, K. L., and W. F. Erbelding. 1981. A photogalvanic cell utilizing the photodissociation of iodine in
solution. Sol. Energy. 27: 139–141.
Suresh Raj, A. M. E., J. Pragasam, F. P. Xavier, and K. S. Nagaraja. 2000. Photoelectrochemical studies on
[MnMoO2(NCS)(Ox)3(H2O)2] (Ox = 8-quinolinol): A novel system for solar energy conversion. Int. J.
Energy Res. 24: 1351–1358.
Suresh, E., J. Pragasam, F. P. Xavier, and K. S. Nagaraja. 1999. Investigation of manganese-molybdenum-
diethyldithiocarbamate complex as a potential system for solar energy conversion. Int. J. Energy Res.
23: 229–233.
Tamilarasan, R., and P. Natarajan. 1981. Photovoltaic conversion by macromolecular thionine films. Nature.
292: 224–225.
Tien, H. T., J. Higgins, and J. Mountz. 1979. Photogalvanovoltaic cells and photovoltaic cells using glassy
carbon electrodes. Trans. J. Br. Ceram. Soc. 203–235.
Tien, T. H., J. M. Mountz, and M. James. 1978. Photo-galvano-voltaic cell: A new approach to the use of solar
energy. Int. J. Energy Res. 2: 197–200.
Titse, A. M., A. M. Timonov, and G. A. Shagisultanova. 1993. Photosensitive chemically modified electrodes
for photogalvanic cells. Coord. Chem. Rev. 125: 43–52.
Tsivadze, A. Y., G. V. Ionova, V. K. Mikhalko, and I. S. Ionova. 2012. Semiconductor-nanoparticle-donor
nanochemical composites: Properties and applications as energy sources of the future. Prot. Metals
Phys. Chem. Surfaces. 48: 1–26.
Tsubomura H., Y. Shimoura, and S. Fujiwara. 1979. Chemical processes and electric power in photogalvanic
cells containing reversible or irreversible reducing agents. J. Phys. Chem. 83: 2103–2106.
Viswanathan, K., and P. Natarajan. 1996. Studies on the photoelectrochemical properties of thionine dye
covalently bound to poly (acrylamidoglycolic acid). J. Photochem. Photobiol. A: Chem. 95: 255–263.
Weber, K., and E. Matijević. 1947. Über photogalvanische Erscheinungen bei organischen Redoxsystemen.
Experientia. 3: 280–281.
Wildes, P. D., D. R. Hobart, N. N. Lichtin, D. E. Hall, and J. A. Eckert. 1977. Sensitization of an iron-thiazine
photogalvanic cell to the blue: An improved match to the insolation spectrum. Sol. Energy. 19: 567–570.
Yadav, S., and C. Lal. 2013. Optimization of performance characteristics of a mixed dye based photogalvanic
cell for efficient solar energy conversion and storage. Energy Conver. Manage. 66: 271–276.
Yamase, T., and S. Ikawa. 1980. Production of hydrogen by photogalvanic cell. Part 2. pH dependence of
alkylammonium molybdate system. Inorg. Chim. Acta. 45: L55–L57.
Yi, X.-Y., L.-Z. Wu, and C.-H. Tung. 2000. Long-lived photoinduced charge separation in [Ru(Bpy)3]2+/viologen
system at nafion membrane-solution interface. J. Phys. Chem. B. 104: 9468–9474.
Zeichner, A., J. R. Goldstein, and G. Stein. 1978. Photogalvanic effect in ferric bromide solutions. J. Phys.
Chem. 82: 1687–1692.
Zhou, R-L., Y. G. Yang, and Y. Y. Han. 1994. The study of photoelectrochemical cells based on chlorophyll a
and safranine T. J. Photochem. Photobiol. A: Chem. 81: 59–63.
Znamensky, D. A., R. G. Yusupov, and B. V. Mislavsky. 1992. Langmuir-Blodgett mono- and multilayers of flu-
orocarbon amphiphilic polymers and their application in photogalvanic metal-insulator-semiconductor
structures. Thin Solid Films. 219: 215–220.
8 An Alternative Fuel
Hydrogen

Neelu Chouhan, Rajesh Kumar Meena, and Ru-Shi Liu

CONTENTS
8.1 Introduction........................................................................................................................... 140
8.2 Hydrogen................................................................................................................................ 141
8.2.1 Chemical Fuel............................................................................................................ 141
8.2.2 Hydrogen as a Sustainable and Clean Energy Source............................................... 142
8.2.3 Hydrogen Fuel: Is It Safe?.......................................................................................... 142
8.3 Hydrogen Production: Past, Present, and Future................................................................... 142
8.4 Concept of Photochemical Hydrogen Generation................................................................. 144
8.5 Technology for Hydrogen Generation................................................................................... 147
8.5.1 Kvaerner Carbon Black and Hydrogen Process......................................................... 148
8.5.2 Biological Production................................................................................................ 148
8.5.3 Electrolysis of Water.................................................................................................. 148
8.5.4 Concentrating Solar Thermal Power......................................................................... 149
8.5.5 Thermal Electrolysis of Water................................................................................... 150
8.5.5.1 Two-Step Cycles......................................................................................... 150
8.5.5.2 Three-Step Cycles: Main Cycles................................................................ 151
8.5.6 Photocatalytic and Photoelectrocatalytic Hydrogen Production............................... 151
8.5.7 Hydrogen as a By-Product of Other Chemical Processes......................................... 152
8.5.7.1 Catalytic Reforming of Oil Refinery.......................................................... 153
8.5.7.2 Chlor-Alkali Plants..................................................................................... 153
8.5.7.3 Hydrocarbon Waste of High-Temperature Fuel Cells................................. 153
8.5.7.4 Waste Biomass of Wine Industries............................................................. 154
8.6 Photocatalytic Hydrogen Generation..................................................................................... 155
8.7 Mechanisms of Photocatalytic Water Splitting..................................................................... 156
8.8 Modification of Photocatalysts.............................................................................................. 157
8.8.1 Photosensitization: Dyes/Quantum Dots................................................................... 157
8.8.2 Reducing Bandgaps through Doping......................................................................... 159
8.8.3 Engineered Solid Solutions........................................................................................ 159
8.8.4 Macromolecular Systems for Water Splitting............................................................ 159
8.8.5 Plasmonic Nanostructures with Surface Plasmon Resonance................................... 161
8.8.6 Nanostructuring of the Photocatalyst........................................................................ 162
8.8.7 Composite Photocatalyst............................................................................................ 163
8.8.8 Photoelectrochemical Water Splitting....................................................................... 163
8.9 Electrochemical Aspect......................................................................................................... 164
8.10 Hydrogen as a Key Solution.................................................................................................. 167
References....................................................................................................................................... 169

139
140 Solar Energy Conversion and Storage

8.1 INTRODUCTION
Hydrogen is the most widely occurring element of the universe. It is available on Earth in multiple
chemical forms such as hydrocarbon, hydrides, prebiotic organic compounds, water, and so on.
All the major common conventional energy sources, such as oil, gas, and coal, are not renewable.
Moreover, their amounts on Earth are being exhausted rapidly. Therefore, the trend is now shifting
toward renewable energy sources. In the last few decades, renewable energy has become an appar-
ent choice over conventional energy. The prime driving forces behind this trend are future carbon
emissions constraints, energy security, downfall of subsisting energy sources (raising prices), result-
ing changes in climate and environment, escalating energy demands of industrial and economic
development, inextricable link between nuclear weapons and nuclear power, abundant and free
renewable energy sources, consumer awareness, financial risk mitigation, flexibility, and resilience.
Renewable energy will no longer be a minor player as it surpasses nuclear energy and other tradi-
tional energy sources. This will occur only on the condition that one happens to crack the code of
sustainable economic growth by reducing energy demands. A futuristic energy scenario is the alter-
native image of how the future might unfold with an appropriate tool to analyze how driving forces
may influence future outcomes and also to assess the associated uncertainties.
Hydrogen is a renewable and clean fuel that offers us the greatest promise of improving our
energy problems. The only combustion product of hydrogen is water vapor without any trace of nox-
ious carbon emissions in the form of carbon dioxide, carbon monoxide, or unburned hydrocarbons.
Now the question arises, how can one get this hydrogen? Because 90% of the material available on
Earth contains hydrogen, it is very tempting to use hydrogen as an energy source (fuel) and energy
carrier. To utilize this resource, pure hydrogen must be obtained in a cost-effective, benign, and reli-
able way. Due to its qualities, hydrogen has attracted much attention from scientists as the fuel of the
future. Currently, the cleavage of natural gas methane is the major source of hydrogen production,
but this requires comparatively harsh conditions of high temperature and pressure and results in the
production of large amounts of greenhouse gas emissions, which have already been linked to ozone
depletion and global warming. Hence, this method is not considered environmentally friendly.
The eco-friendly production of hydrogen in its purest form is really a difficult task. Environmentally
benign hydrogen can be produced by the cleavage of water by using renewable energy sources
such as solar energy. Photocatalytic or photoelectrochemical water-splitting techniques are used
for hydrogen production. In this context, one must contend with the difficulty of, the low density
of the energy when considering the use of solar energy. A large area is required in order to har-
vest a reasonable amount of solar energy. Although there are many methods of splitting water—
thermochemically, biophotolytically, mechanocatalytically, plasmolytically, magnetolytically, and
radiolytically—photocatalytic water splitting will be advantageous for the large-scale application of
solar hydrogen production because of its simplicity. It is the most promising route, and it is

1
H 2O + hν + Photocatalyst → O 2 + H 2 (8.1)
2
Nowadays, hydrogen is primarily used as a fuel in spaceships and space shuttles to lift them into
space and supply electricity to all the systems during the flight. Speedy development in related tech-
nologies, such as hydrogen fuel cells, made it possible to use this gas not only in space shuttles but
also in ordinary road vehicles, cars and buses, making them much cleaner without posing any threats
to the environment. Hydrogen is also considered as a potential energy carrier, just like electricity,
which is the primary energy carrier of the world and is utilized for transporting persons and goods
from one place to another. On mixing hydrogen with oxygen in fuel cells, electricity is generated,
with water being the only by-product and without any trace of harmful carbon emission. It represents
an absolutely green form of energy that can be directly used as an energy source like oil, gas, or coal.
This reaction is extraordinarily slow at ambient temperature and moderately accelerated in the
presence of a catalyst, such as platinum, or an electric spark. The production of hydrogen will allow
Hydrogen 141

for energy independence. The benefits of hydrogen vehicles are clear and more realistic in fifth-
generation vehicles that feature new design possibilities (stack of 200 cells size equal to a home
PC), quieter operation (no moving parts in the fuel cell stack), eco-friendly operation (zero carbon
emission), possibility of interchangeable bodies (skateboard-like chassis), no engine compartment,
larger cabin 82 hp engine 0–62 mph quick start (in 16 s), and high speed. The Honda FCX concept
car, with a maximum speed of 93 mph, 80 hp, and 201 lb-ft torque (Burns et al. 2002), is about to
come on the market, and it is hoped that the dream to drive a hydrogen car will be fulfilled by 2020.
The development and use of hydrogen fuel cells mean a decreased need for Middle Eastern oil or low
imports of hydrogen by the United States or a non-oil-producing country. It can also lead to possible eco-
nomic collapse in main oil-exporting nations. Hydrogen fuel cells can be used to power third world coun-
tries. In underdeveloped countries, the governments would not have to import oil. Solar or wind power
collectors could produce energy, which would make hydrogen. Hydrogen fuel cells will drastically change
our cars and the ways we heat and power our homes and businesses. This will lead to positive and negative
changes in the lifestyles of laymen and political scenarios of the Middle East and third world countries.

8.2 HYDROGEN
8.2.1  Chemical Fuel
Hydrogen is the simplest element in the universe with one proton and one electron. It is the most
abundant element on Earth. Fortunately or unfortunately, it cannot remain alone. At normal tem-
perature, it shows low reactivity unless it has been activated by an appropriate catalyst, but at high
temperatures, it becomes highly reactive. Atomic hydrogen is treated as a powerful reductive agent at
room temperature and produces hydrogen peroxide, H2O2, with oxygen. It reacts with the oxides and
chlorides of metals, like silver, copper, lead, bismuth, mercury, and so on, to produce free metals. It
reduces some salts to their metallic state, like nitrates, nitrites, and sodium and potassium cyanide. It
reacts with a number of elements, metals, and nonmetals, to produce hydrides, like NAH, KH, H2S,
and PH3. Similarly, it reacts with organic compounds to form a complex mixture of products (e.g.,
with ethylene, C2H4, the products are ethane, C2H6, and butane, C4H10).
Hydrogen makes up almost 95% of the visible matter (mass) on the Earth. It is an odorless, invis-
ible, nontoxic and nonpoisonous, smoke free, and highly buoyant (lighter than air) gas that rises and
diffuses when leaked. The H2 gas can be most easily identified by the thermal wave, which produces
low radiant heat. With the widest range of flammability (H2: 4%–75%, gasoline: 1%–7.6%), low auto-
ignition temperature (400°C), and high octane rating (130; gasoline: 87–93), hydrogen has the highest
heating value among all available fuels at nearly 52,000 Btu lb–1 (British thermal units per pound
mass). Its flammability range allows for lean mixtures with better fuel economy and lower combustion
temperatures than gasoline; hence, hydrogen engines perform more efficiently than gasoline engines.
It possesses a compression ratio higher than gasoline and lower than diesel, which makes it beneficial
when used as a fuel for internal combustion engines (ICEs), because lean mixtures increase efficiency
(diesel engines > hydrogen engines > gasoline [petrol] engines), especially at low power and at engine
idle, and lower combustion temperatures can help to suppress the amount of emissions of nitrogen
oxides, which can still occur in ICEs fueled by hydrogen. Moreover, in terms of efficiency, 57 mil-
lion metric tons of hydrogen is equal to about 170 million tons of oil equivalent. The growth rate of
hydrogen production is around 10% per year. These superb properties of hydrogen strongly support
hydrogen being considered as a green and clean fuel. Some other well-known alternative fuels are bio-
diesel, bioalcohol (methanol, ethanol, butanol), chemically stored electricity (batteries and fuel cells),
nonfossil methane, nonfossil natural gas, vegetable oil, propane, and other biomass sources. The main
purpose of any fuel is to store energy, which should be in a stable form and can be easily transported
to the place of production. Almost all fuels are chemical fuels. The user employs this fuel to generate
heat or perform mechanical work, such as powering an engine. It may also be used to generate elec-
tricity, which is then used for heating, lighting, or other purposes. Furthermore, high diffusivity, low
142 Solar Energy Conversion and Storage

viscosity, unique chemical nature, combustibility, and electrochemical properties are the characteris-
tics that also make the hydrogen a different or better fuel than other gases.

8.2.2 Hydrogen as a Sustainable and Clean Energy Source


Renewable energy sources for electricity constitute a diverse group by including the wind, solar, tidal,
and wave energy to hydro, geothermal, and biomass-based power generation. Apart from hydro power
in the few places where it is plentiful, none of these is suitable, intrinsically or economically, for large-
scale power generation, where a continuous and reliable supply of energy is needed. Growing use will,
however, be made of renewable energy sources in the years ahead, although their role is limited by their
intermittent nature. Another issue associated with renewable energy source use is economic feasibility,
which becomes crucial when an intermittent supply of electricity is demanded at a small scale. In the
Organisation for Economic Co-operation and Development (OECD) countries, about 2% of electricity
is harnessed from renewable sources other than hydro, and this is expected to increase to 4% by 2015.
Hydrogen production is a large and growing industry and is useful as a compact energy source
in fuel cells and batteries. Hydrogen can be produced using renewable energy sources like wind,
solar, waves, hydro, and biomass. The modest form of renewable energy can be stored in water as
hydrogen, produced by splitting water by means of electrolysis, and used for electricity. Relying
only on renewable energy may solve the threats of oil depletion and pollution in the present energy
system. This also makes it possible for everybody to produce their own energy, creating more politi-
cal stability and benefits to all of us. After the production, hydrogen functions as an energy carrier
that can be used to supply energy wherever it is needed.
Hydrogen and fuel cells are comparatively new and different energy systems that either use
energy directly as electricity or store it in the form of hydrogen for use in transportation or store for
the time when the sun is not shining or the wind is not blowing. The fuel cells are used to convert
the hydrogen into energy, where hydrogen and oxygen (air) react and produce water as the only
emission. The reaction creates electricity and heat that can be used in various applications. Fuel
cells are versatile in their application and can be used in many devices that need energy, including
mobile phones, cars, buses, and even heat and power plants. Fuel cell technology is the next innova-
tive development that will bring progress and prosperity to society and will remarkably impact our
lives, as significantly as did the steam engine and the combustion engine.

8.2.3 Hydrogen Fuel: Is It Safe?


The superiority of hydrogen fuel over other fuels is an established fact. All fuels are hazardous, but
hydrogen is comparably or less so because of its nonpoisonous and highly buoyant nature. Spills
present very little danger as it rapidly evaporates and disperse. However, it can still displace oxygen
in confined spaces. There is a need to handle it with care, because it can burn with a clear flame and
without smell, cannot sear from a distance, burns first with no smoke, does not puddle, and is hard
to make explode in free air (22 times less explosive power). Beyond the Hindenburg myth (1937)
(McAlister 1999), until now nobody was killed by hydrogen fire, which is completely unrelated to
hydrogen bombs. For a vehicle, it can be stored onboard in either pressurized gas or liquid form in
composite tanks that utilize carbon fiber for additional safety.
Hence, one can say that hydrogen is a safe energy source (fuel), if one can handle it carefully.

8.3  HYDROGEN PRODUCTION: PAST, PRESENT, AND FUTURE


There are some inherent advantages of hydrocarbons (natural gas and petroleum), including feedstock
availability, cost competitiveness, convenience of storage and distribution, purity of the product, and
relatively high H/C ratio. Scientists have been highly engaged in research in the field of hydrogen
production. The current world’s major resources for hydrogen generation are coal (18%), natural gas
Hydrogen 143

Electrolysis

Electrolysis 4%
Coal Coal 18%
Oil
Natural gas 48%
Oil 30%

Natural gas

FIGURE 8.1  Sources of hydrogen generation from natural


Steam methane reforming
gases, oil, coal, and electrolysis of water.

(40%), petroleum (38%), and electrolysis of water (4%) that consume approximately 50 kW h–1 of elec-
tricity for per kilogram of hydrogen production (Hairston 1996) (Figure 8.1). Most of the industrial
hydrogen consumers are chemical process industries (73.9%), metallurgical industries (2.7%), food
(3.6%), electronics (8.1%), petroleum, pharmaceuticals (11.7%), and other industries (Hairston 1996).
Scientists have found dramatic growth in the demand for renewable energy over the past few
decades, which reflects the “take-off” phase of the renewables market along with energy supply
security, autonomy, resilience, jobs, industrial development, financial profit, portfolio risk mitiga-
tion, price risks of fossil fuels, rural energy access, climate change, environmental sustainability,
and reduction in nuclear accidents/wastes. Improvement in the economy for large-scale manufactur-
ing, radical improvements in technology, advances in performance, and reductions in the costs of
renewable energy (wind, hydropower, water, geothermal, and biomass), categorize hydrogen as the
most important source of energy. Onshore wind power is closest to commercial maturity. Moreover,
offshore wind power is more expensive than onshore, but it has a large (although uncertain because
of fluctuating density) potential for cost reductions, not just for turbines, but also for logistics, long-
term operations, and maintenance costs. Remarkable reduction in the cost of solar photovoltaics
(PV) has been seen in recent years. Solar energy input in a year can replace the total world storage
of conventional fuel and nuclear fuel. Therefore, at present, annual global investment in power gen-
eration is flowing to renewable energy instead of fossil fuels and nuclear. The growth of renewable
energy worldwide began in the 1990s and has greatly accelerated in the 2000s.
Many have credited this growth to the proliferation of supportive government policies, rising costs of
conventional energy, dramatic reductions in renewable energy technology costs, and economies of scale
in manufacturing. For these major reasons, dynamic growth has accelerated over the past decades, and
past projections about renewable energy have fallen short. The International Energy Agency (IEA) in
2000 projected 34 GW (gigawatts) of wind power globally by 2010, while the actual level reached was
200 GW. In 1996, the World Bank projected 9 GW of wind power and 0.5 GW of solar PV in China
by 2020, while the actual levels reached in 2011 were 62 GW of wind power and 3 GW of solar PV
(Hairston 1996). The history of energy scenarios is full of similar projections for renewable energy that
proved to be too low by a factor of 10, or were achieved a decade earlier than expected.
Projections show continued cost reductions, many possible technology advances, and competi-
tiveness with retail electricity prices without subsidies occurring in many jurisdictions soon and
in many more places around the world by 2020. Concentrated solar power (CSP) has a large cost
reduction potential, with future opportunities for bulk power supply, for dedicated applications such
as industrial heat supply and desalination, and power grid balancing using multihour and multiday
embedded heat storage. Another promising advance in these series is the first generation of biofuels.
Many projections show large future markets for advanced biofuels from agricultural and forestry
wastes, and from crops grown on unproductive lands. A wide variety of new approaches for use
of biomass are also projected, such as growing international commodity markets for wood pellets
and bio-heating oil, greater use of biogas in a variety of applications, new types of biorefineries in
agriculture and forestry, and greater use of biomass in heat supply.
144 Solar Energy Conversion and Storage

At least 30 countries already have shares of renewable energy more than 20%. Some 120 coun-
tries have various types of policy targets for long-term shares of renewable energy (e.g., 20% target
for the European Union by 2020). Some countries have long-term policy targets that will put them
in the high renewables domain by 2030 or 2050, such as Denmark (100%) and Germany (60%). A
diverse group of 20 countries outside of Europe target 10%–50% energy shares in the 2020–2030
time frame (Ogden 1999).
Renewable markets will become even broader based in a larger number of countries beyond
2020, as developing countries increasingly share leadership. Unique opportunities for renewable
energy exist in future development that includes new electric power infrastructure, diesel genera-
tor replacement, new settlements, new power-market rules, regional cooperation frameworks, local
manufacturing, and rural (off-grid) energy services. Furthermore, the predicted decrease in conven-
tional sources as well as the interrupted imported oil supply due to political instability in the Middle
East, will result in increased petroleum prices.
Hydrogen fuel is rich in energy and is basically proposed as an alternative fuel that has the poten-
tial to dramatically decrease the energy consumption of vehicles. On using it in a fuel cell, electric-
ity and heat are produced by burning H2 and O2. The by-products in this process are H2O and, of
course, small amounts of nitrogen oxide. Because the consumption of hydrogen as a fuel produces
nonpollutant flues, many experts believe that hydrogen fuel could be an eco-friendly alternative to
our current energy sources, which can be commercialized in the future on a large scale. Hydrogen
was first commercially used as a fuel in a fuel cell in the 1970s for launching NASA’s space shuttles
and other space rockets. Since then, hydrogen fuel cells have also been used in the production of
electricity for spacecrafts while in orbit. Astonishingly, the water produced from burning H2 and O2
during long travels in space can even be used for drinking purposes for the crew, which proves that
hydrogen fuel is indeed an eco-friendly energy source.
Fuel cells are often compared to batteries. Both produce energy by conversion of a chemical reac-
tion into usable electric power. However, the fuel cell will produce electricity as long as fuel (hydrogen)
is supplied, never losing its charge. A fuel cell is a promising technology for use as a source of heat
and electricity for buildings and electric motors propelling vehicles. Fuel cells operate best on pure
hydrogen. Fuels like natural gas, methanol, or even gasoline can be reformed to produce the hydrogen
required for fuel cells. Some fuel cells can be fueled directly with methanol, without using a reformer.
Outstanding fuel efficiency of hydrogen is always welcome, but associated technological challenges
owing to its lightness (i.e., safety in production, storage, transportation, distribution, and end use) are
making its path difficult (Ogden 1999; Szklo and Schaffer 2007; Scholz 1993). Because hydrogen is
a low-density gas, imperfect packing and careless refilling into a cylinder may cause its leakage into
the atmosphere. It was estimated by Caltech Research that atmospheric emission will grow from 60 to
120 trillion (four to eight times more hydrogen than the normal human action release into air) a year
by replacing fossil fuel with hydrogen globally. Although the figures remain uncertain until scientists
gain more understanding of the effects of using hydrogen as an energy source, some approximations
have been made of a 10% decrease in ozone concentration in the planet’s atmosphere caused by the
inevitable leaks. That was attributed to moisture buildup from hydrogen combining with atmospheric
oxygen. As a consequence, the upper atmosphere cools, causing indirect destruction of the ozone.
Hydrogen has the potential to be used as an energy carrier in electricity production. It can be
considered as an energy carrier, which moves and may deliver energy in a usable form to consumers
as and when required. Other renewable energy sources, such as the Sun and wind, cannot produce
energy all of the time. But they could produce electric energy and hydrogen, which can be stored
for future needs. Hydrogen can also be transported (like electricity) to locations where it is needed.

8.4  CONCEPT OF PHOTOCHEMICAL HYDROGEN GENERATION


Among the variety of available hydrogen-containing compounds on Earth, water is a carbon-free
abundant source of hydrogen. It is transparent by appearance and categorized as almost unbreakable
Hydrogen 145

(stable) under normal circumstances. Moreover, due to the low autoionization (Kw = 1.0 × 10−14),
pure water falls into the category of insulator at room temperature that poorly conducts current,
0.055 µS⋅cm−1. Therefore, very large potential is required to increase the autoionization of water.
Furthermore, kinetic and thermodynamic decomposition of pure water into hydrogen and oxy-
gen is not suitable at standard temperature and pressure due to large positive Gibbs free energy
(ΔG  =  237  kJ mol−1); thus, the backward reaction (recombination of hydrogen and oxygen into
water) easily proceeds (Chouhan and Liu 2011). Therefore, activation energy, ion mobility (dif-
fusion) and concentration, wire resistance, surface hindrance including bubble formation (causes
electrode area blockage), and entropy require a greater applied potential to overcome these factors.
The following oxidation and reduction reactions are involved in water splitting (Equations 8.2 and
8.3) at the anode and cathode:

Anode: 2H 2O (l ) → O 2 ( g ) + 4H + ( aq ) + 4e− Eo,ox = –1.23 V ( Eo,red = 1.23) (8.2)

Cathode: 2H + ( aq ) + 2e− → H 2 ( g ) Eo,red = 0.00 V (8.3)

Thus, the standard potential of the water electrolysis cell is –1.23 V at 25°C at pH = 0 (H+ = 1.0 M)
and pH = 7 (H+ = 1.0 × 10−7 M), based on the Nernst equation.
A well-known example of photocatalysis occurring in nature is photosynthesis by plants, where
chlorophyll serves as the photocatalyst. Photocatalytic splitting of water mimics the photosynthe-
sis process (Figure 8.2) as both work on the same principle—that is, under irradiation of light an
electron is excited from the valence band into the conduction band to result in the formation of an
electron (e−)–hole (h+) pair in a photocatalyst. These e− and h+ reduce and oxidize any substrate on
the surface of the photocatalyst, respectively; if they recombine, then there will be no net chemical
reaction. The original structure (or chemical composition) of the photocatalyst remains unchanged
if an equal number of e− and h+ are consumed for chemical reaction and/or recombination.
Because of its abundance and free availability, solar energy has become the most obvious and
preferred option for water breaking, which needs photons of the light wavelength shorter than
1008 nm, which corresponds to 1.23 eV. Fortunately, water is transparent; hence, automatically, it
is unable to break down in ordinary conditions. Theoretically, if the energy gap of samples is kept
at 1.23 V, then the maximum value of solar energy conversion efficiency will be about 32%. But
practically, the conversion efficiency should be in the range of 10%–15%, because of the various
overpotentials of the system. The bandgap is kept between 1.8 and 2.5 eV to avoid recombination
(2.46 eV). Therefore, the material issue became vital for the water-splitting process, but the nature
of electrolytes used and fabrication technology are also important concerns for their industrial
applications.
Several terms have been adopted to describe the efficiency for converting solar energy, namely,
the incident photon-to-current efficiency (IPCE), absorption photon-to-current efficiency (APCE),
solar-to-hydrogen (STH) conversion efficiency, and quantum efficiency (QE) (Murphy et al. 2006).

Photosynthesis CO2 + H2O Sugar + O2


H2 + O2

∆G0 = +237.2 kJ mol-1


Photocatalysis H2O

FIGURE 8.2  Analogy of photosynthesis with photocatalytic water splitting.


146 Solar Energy Conversion and Storage

Percentage IPCE is calculated using Equation 8.4, which includes photocurrent density, wavelength
of light, and irradiance:

IPCE (%) =
(1240 × Photocurrent density) (8.4)
(λ × Irradiance)
The APCE is obtained by dividing the IPCE by the fraction of incident photons absorbed at each
wavelength by light-harvesting efficiency (LHE), calculated through the absorbance A (Equations 8.5
and 8.6). APCE is usually used to characterize the photoresponse efficiency of a photoelectrode
material under an applied voltage:

IPCE
APCE = (8.5)
LHE (λ)

⎛α ⎞
LHE(λ) = TTCO(λ)⋅ ⎜ dye ⎟ × 1− eα (8.6)
( )
⎝ α film ⎠

where:
TTCO is the transmittance of the transparent conductive oxide
αfilm is the absorption coefficient of the entire film
αdye is the absorption coefficient due to the dye molecules

This is the simplest approach for calculating APCE, where second-order reflectance terms are not
considered. The LHE is estimated by using the information about injection and collection efficien-
cies (Equation 8.6). The solar-to-hydrogen conversion efficiency STH of the water-splitting reaction
can be determined using Equation 8.7:

100
STH (%) = j p (VWS – VBias ) × (8.7)
eEs

where:
jp is the photocurrent density (mA cm−2) produced per unit irradiated area
VWS (1.23 eV) is the water-splitting potential per electron
VBias is the bias voltage applied between the working and the counterelectrodes
ES is the photon flux
e is the electronic charge

According to Equation 8.7, every electron contributing to the current produces half an H2 molecule.
The photoconversion efficiency (η) is a percentage conversion of light energy into chemical energy
in the presence of applied external voltage, and it can be calculated using Equations 8.8 and 8.9:

η (%) = ( Total power output − Electrical power output ) ×100 (8.8)

⎡ ( E0,rev – Eapp ) ⎤
= jp ⎢ ⎥ ×100 (8.9)
⎢⎣ I0 ⎥⎦

where:
Eapp = Emeas – Eaoc
E 0rev is the standard reversible redox potential of the water and is equal to the value 1.23 versus
NHE
Hydrogen 147

Emeas is the electrode potential (vs. Ag/AgCl or saturated calomel electrode [SCE]) of the work-
ing electrode at which photocurrent was measured under light illumination
Eaoc is the electrode potential (vs. Ag/AgCl) of the same working electrode at open-circuit con-
dition under the same illumination and in the same electrolyte solution
I0 is the intensity of incident light (mW cm−2)

Quantum efficiency (QE) represents the characteristic photon conversion of photoactive films. It is
defined as the percentage of generated electrons and incident photons, while the photoactive films
are irradiated under a specific wavelength (Equation 8.10). It is noted that QE neglects the energy
loss of solar irradiance and the chemical conversion efficiency. Therefore, it is used to qualify the
photoactive films but not to represent the water-splitting reaction conversion efficiency. The appar-
ent quantum efficiency was measured under the xenon arc lamp or four 420 nm-LEDs (3 W), used
to trigger the photocatalytic reaction:

⎛ Number of evolved H 2 molecules × 2 ⎞


QE (%) = ⎜ ⎟ ×100 (8.10)
⎝ Number of incident photons ⎠

8.5  TECHNOLOGY FOR HYDROGEN GENERATION


Progress of the human race revolved around the axis of fulfillment of our energy demand, which can
facilitate the development of technologies and the productivity that is necessary for a higher stan-
dard of living. Therefore, energy is as vital as raw materials, capital, and labor, for growth. Fossil
fuels are the most commonly available form of the energy and transportation fuel, but they are not
ideal fuels because of the following:

• The combustion of fossil fuels, such as natural gas, coal, and petroleum, emits carbon
dioxide and other pollutants, which is one of the major components of greenhouse gases
that cause notorious climate changes.
• The supply of hydrocarbon resources in the world is limited, and the demand for hydro-
carbon fuels is increasing day by day, particularly in China, India, and other developing
countries. Various biological and nonbiological processes have taken place for millions of
years, resulting in the storage of energy in mineral organic compounds or fossil fuels such
as coal, petroleum, and natural gas. If their rates of production do not match the rate of
consumption, that will lead to them being quickly exhausted.
• The supply of fossil fuels is powered by a limited number of nations. As a result, a significant
amount of time and money will be spent for the relocation and distribution of these fuels. It
was also seen that centralized power will cause the gasoline resources war among the nations.
• World transportation is primarily fueled by petroleum fuel (hydrocarbon economy).
Therefore, it will become necessary for us to develop a sustainable energy source that can
be easily produced at low cost and, of course, in an environmentally friendly way. Several
alternative renewable energies have been developed to replace or reduce the use of fos-
sil fuels. Alternative energies have lower carbon emissions as compared to conventional
energy sources. These energy sources include wind, hydropower, solar, geothermal, and
so on, which can be utilized to generate clean and efficient fuel (i.e., hydrogen, which is
abundantly available in nature in various forms). The Earth’s crust contains only 0.15%
natural H2 and 1.0% in atmosphere as free hydrogen. Hydrogen is a major constituent (0.5
ppm) of the most abundantly available resource in substituted form (98%)—that is, water
(70%) and biomass (14%). Presently, hydrogen is mainly produced from fossil sources.
This stimulated great interest in alternative, cheaper means of hydrogen production.
Proponents of a world-scale hydrogen economy argue that hydrogen can be an environ-
mentally cleaner source of energy to end users, particularly in transportation applications,
148 Solar Energy Conversion and Storage

without the release of pollutants (such as particulate matter) or carbon dioxide at the point
of end use. An analysis asserted that most of the hydrogen supply chain pathways would
release significantly less carbon dioxide into the atmosphere than would gasoline used in
hybrid electric vehicles. As a result, a significant reduction in carbon dioxide emissions
would be possible, if carbon capture or carbon sequestration methods were utilized at the
site of energy or hydrogen production.

Centralized production of hydrogen requires a large capital investment for the construction of the
infrastructure and distribution to filling stations that may put a fleet of light-duty fuel cell vehicles
on the road. Furthermore, the technological challenge of providing safe, energy-dense storage of the
hydrogen on board a vehicle must be overcome to provide sufficient distance between fill-ups. Some
chemical/technical methods involved in hydrogen production are discussed here.

8.5.1  Kvaerner Carbon Black and Hydrogen Process


A Norwegian company (Kvaerner) developed an endothermic method for the production of hydro-
gen from hydrocarbons (CnHm), such as methane, natural gas, and biogas, in the 1980s. However, it
was first commercially exploited in 1999. It is an efficient mode of complete transformation of the
hydrocarbon/methane to pure carbon (40%), heat (2%), and hydrogen (48%), using a plasma con-
verter at around 1600°C in comparison to other reformation methods such as steam reforming and
partial oxidation the natural gas (Naess et al. 2009):

m
Cn H m → nC + + Heat (8.11)
2H 2

8.5.2 Biological Production
Biohydrogen is manifested by a diverse group of living objects, such as bacteria, plants, multi-
enzyme systems on organic substrate, and so on, involving steps similar to anaerobic conversion
through photosynthesis, fermentation, and microbial electrolysis of cells. It was discovered in the
1990s that if an algae is deprived of sulfur, it will switch from the production of oxygen (normal
photosynthesis) to the production of hydrogen. Biological hydrogen production provides a possibil-
ity of being renewable and carbon neutral as a method of H2 production, and it can also be produced
in an algae bioreactor, using sulfur-deprived algae.
The process involves bacteria feeding on hydrocarbons and excreting hydrogen and CO2. The
CO2 can be sequestered successfully by several methods that result in the production of pure
hydrogen gas. A prototype hydrogen bioreactor using waste as a feedstock is already in operation.
Biological hydrogen can also be generated by the fermentation process in dark and light. In dark,
fermentation reactions do not require light energy, but the photofermentation proceeds in the pres-
ence of light, and it can be employed in the conversion of small molecular fatty acids into hydrogen
by using Rhodobacter sphaeroides SH2C (Tao et al. 2007). They are capable of constant production
of hydrogen from organic compounds throughout the day and night.
Another important means of biological hydrogen production is biocatalyzed electrolysis (i.e.,
electrolysis using microbes). Hydrogen is generated after running through the microbial fuel cell
with biocatalyzed electrolysis and a variety of aquatic plants, such as reed sweet grass, cord grass,
rice, tomatoes, lupines, and algae (Rosenbaum et al. 2010).

8.5.3 Electrolysis of Water
At ambient conditions, pure water cannot dissociate due to a very poor degree of self-ionization
(Kw  =  1.0  ×  10−14). Furthermore, the decomposition of pure water into hydrogen and oxygen is
Hydrogen 149

thermodynamically not favorable at standard temperature and pressure (Gibb’s free energy
ΔG 0 = 237 kJ mol−1; 2.46 eV per molecule) (Equation 8.12). Therefore, it is necessary to associate
a large potential to induce electrolysis in water. This additional potential is known as overpotential.
When this reaction is about to complete after applying potential, it is required to overcome certain
factors (i.e., activation energy, ion mobility [diffusion] and concentration, wire resistance, surface
hindrance including bubble formation [causes electrode area blockage], entropy, etc.):

Overall: 2H 2O (l ) + 4e− → O 2 ( g ) + 2H 2 ( g ) + 4e− E = –1.23 V at 25°C; pH = 0 and 7 (8.12)

Thus, the standard potential of the water electrolysis cell is –1.23 V at 25°C, based on the Nernst
equation. Water-soluble electrolytes or electrocatalysts are added to improve the conductivity of the
water along with the applied potential. Salts of cations having electrode potential lower than H+,
such as Li+, K+, Na+, Rb+, Cs+, Mg2+, Sr2+, Ca2+, and Ba2+, are used as the electrolyte (Pauling 1970).
A solid polymer electrolyte such as Nafion (1.5 V) can also be utilized. The electrolyte disassociates
into cations and anions in water. Afterward, these cations and anions will rush toward the cathode
and anode to neutralize the negatively charged OH− and positively charged H+ ions, at respective
electrodes. This phenomenon allows the continued flow of electricity (Pauling 1970). The electro-
catalysts, such as platinum alloys, molybdenum sulfide (Kibsgaard et al. 2014), graphene quantum
dots (Fei et al. 2014), carbon nanotubes, and perovskite (Luo et al. 2014), are suitable for anode
material utilized for the oxidation. Similarly, a two-electron reaction to produce hydrogen at the
cathode can be electrocatalyzed with almost no overpotential by platinum and less overpotential
with other materials.
Deiman and van Troostwijk were the first ones, who used gold electrodes for electricity genera-
tion from water. Later on, Volta invented the Voltaic pile for electrolysis of water. In this series,
Gramme invented the Gramme machine for hydrogen production through the electrolysis of water,
while Lachinov developed a method for industrial hydrogen production through the electrolysis of
water. With the appropriate electrodes and electrolyte, oxygen will collect at the positively charged
electrode (anode), and hydrogen will collect at the negatively charged electrode (cathode).
The efficiency of modern hydrogen generators is measured by power consumed per standard
volume of hydrogen (MJ  m−3), assuming standard temperature and pressure of the H2. A 100%
efficient electrolyzer would consume 11.7 MJ m−3. Lower is the actual power used, higher is the effi-
ciency. There are two main technologies available in the market, cheaper alkaline (nickel catalysts,
60%–75%) and more expensive proton exchange membrane (PEM) (platinum group metal catalysts,
65%–90%) electrolyzers. Then 1 kg hydrogen (which has a specific energy of 143 MJ kg−1, about
40 kWh kg−1) generation through water electrolysis requires 50–79 kWh. It is 3–10 times the price
of hydrogen produced from steam reformation of natural gas (NREL 2009). The price difference
is due to the efficiency of direct conversion of fossil fuels to produce hydrogen, rather than burning
fuel to produce electricity. Hydrogen from natural gas, used to replace, for example, gasoline, emits
more CO2 than the gasoline it would replace, and so this process is of no help in reducing green-
house gases (Crabtree et al. 2004). Hydrogen can be made via high-pressure electrolysis or low-
pressure electrolysis of water. The current best processes have an efficiency of 50%–80% (Zittel
and Wurster 1996).

8.5.4  Concentrating Solar Thermal Power


CSP systems generate solar power by using solar concentrators (namely, parabolic trough, enclosed
trough, dish Stirlings, concentrated linear Fresnel reflector, solar power tower, etc.) that concentrate
a large area of sunlight onto a small area. The concentrated light is converted into heat, which drives
a heat engine (usually a steam turbine) connected to an electrical power generator. It generates elec-
tricity, which is used to break water (Figure 8.3).
150 Solar Energy Conversion and Storage

H2
Sun H2O

Heat O2
engine Electricity
generator Electrolyzer

Concentrating
solar heat collector

FIGURE 8.3  Solar thermal hydrogen production system.

CSP is an emerging technology in the field of electricity generation through thermal power,
which turns sunlight either directly into electricity or first thermally breaks water into H2 and O2
and then into electricity. CSP requires clear skies and strong sunlight, which has virtually unlimited
power. The suitable sites for CSP technologies are found between 15° and 40° parallels and occa-
sionally at higher latitudes.
This is unconventional route for cost-effective hydrogen production from solar energy without
emitting carbon dioxide. Solarothermal electrolysis of water was triggered by solar-to-thermal con-
version followed by two steps: (1) thermal to electrical and (2) electrical to chemical conversions.
Very high temperatures (2500 K) are required to dissociate water into hydrogen and oxygen.
New CSP stations present a healthy and economic competition with fossil fuels. This technology
works efficiently under the high solar radiation zone. It was indicated that the metal nitrates (cal-
cium, potassium, sodium)-based thermal storage systems will make the CSP plants highly effective.

8.5.5 Thermal Electrolysis of Water


Solar energy could be used for cost-effective hydrogen production via water splitting without emit-
ting carbon dioxide. A very high temperature (2500 K) was required for dissociation of the water
into hydrogen and oxygen. Plenty of catalysts are available to be used in these processes at compara-
tively feasible temperatures. There are more than 352 thermochemical cycles that can be used for
water splitting, and about a dozen of these cycles, such as the iron oxide cycle, cerium (IV) oxide–
cerium (III) oxide cycle, zinc–zinc oxide cycle, sulfur–iodine cycle, copper–chlorine cycle, and
hybrid sulfur cycle, are actively being researched or in trail phases to produce hydrogen and oxygen
from water and heat without using electricity (Kolb et al. 2007; Chen et al. 2007; Perret et al. 2007).
Abanades and Flamant (2006) have studied a single-step thermal decomposition of methane without
using catalyst that produces hydrogen-rich gas and high-grade carbon black from concentrated solar
energy and methane. A solar concentrator (2 m diameter) with 1500 K mean temperature at nozzle has
shown a thermal-to-chemical conversion efficiency in the range of 2.6%–98%). Charvin et al. (2008)
made a process analysis of ZnO/Zn, Fe2O3/FeO, and Fe2O3/Fe2O4 thermocycles and found these to be
potentially efficient (two- or three-step routes) for large-scaled hydrogen production by concentrated
solar thermal energy that operated at a temperature up to 2000 K with real efficiency reported as 25.2%,
28.4%, and 22.6%, respectively, for water splitting. Perkins and Weimer (2004) studied the solar-to-
thermal water splitting efficiency of 16%–21% using a three-step Mn2O3/MnO cycle.

8.5.5.1  Two-Step Cycles

Fe 3O4
η = 17.4%, cost = 7.86 – 14.75$ kg −1 H 2
( )
FeO
Hydrogen 151

(i) 3 FeO + H 2O (g ) → Fe 3O 4 + H 2 → 850 K (8.13)

1
(ii) Fe 3O 4 → 3FeO + O 2 (g ) →1900 K (8.14)
2
Cd
(η = 48.3%, cost = 4.5$ kg −1 H 2 )
CdO

(i) Cd (s, l) + H 2O → CdO (s) + H 2 (g ) → 298 – 573 K (8.15)


1
(ii) CdO (1) → Cd (g ) + O 2 (g ) →1773 K (8.16)
2
CeO2
(η = NA cost = NA)
Ce 2O3
1
(i) 2 CeO 2 (s) → Ce 2O 3 (s) + O 2 (g ) → 2300 K (8.17)
2
(ii) Ce 2O 3 (s) + H 2O (g ) → 2CeO 2 (s) + H 2 (g ) → 700 K (8.18)

8.5.5.2  Three-Step Cycles: Main Cycles

Fe 2O3
(η = 18.6%, cost = 8.4$ kg −1 H 2 )
Fe 3O4
1
(i) 3Fe 2O 3 → 2Fe 3O 4 + O 2 →1600 K (8.19)
2
(ii) 2Fe 3O 4 + 6KOH → 6KFeO 2 + H 2 + 2H 2O → 673 K (8.20)

(iii) 6KFeO 2 + 3H 2O → 3Fe 2O 3 + 6KOH → 373 K (8.21)

Mn2O3
(η = 16 – 21%, cost = NA)
MnO
1
(i) Mn 2O 3 → 2MnO + O 2 →1873 K (8.22)
2
(ii) 2MnO + 2NaOH → H 2 + 2NaMnO 2 → 900 K (8.23)

(iii) 2NaMnO 2 + H 2O → Mn 2O 3 + 2NaOH → 373 K (8.24)

These processes can be more efficient than high-temperature electrolysis, typical in the range
from 35% to 49% LHV efficiency. Although several thermochemical methods of hydrogen genera-
tion have been demonstrated in laboratories, none has been utilized at production levels because of
high energy requirements and production costs. Chemical-solar routes of hydrogen production with
different combinations are represented in Figure 8.4.

8.5.6 Photocatalytic and Photoelectrocatalytic Hydrogen Production


Hydrogen and electricity play a crucial role as energy carriers and fuel in setting our future energy
endeavors or economics. It is known that the theoretical potential for the water-splitting process is
1.23 eV per H2O molecule. This energy corresponds to the wavelength 1008 nm that makes almost
70% of the solar irradiated photon eligible for water splitting. But practically, our needs are for
152 Solar Energy Conversion and Storage

Photocatalytic

Heat CST Electrolysis


H2
PV Electrolysis

Heat Solar thermochemical

Photobiological

FIGURE 8.4  Technical routes for solar to hydrogen conversion through photocatalytic, concentrating solar
thermal (CST) and photovoltaic and solar thermochemical (STCH) methods.

higher energy (1.8–2.6 eV) than the minimum energy 1.23 eV, which is required for compensating
for intrinsic energy losses associated with the redox reactions, proceeding on the surface of the
photocatalyst (Dincer and Joshi 2013). Photocatalytic materials require a minimum of sunlight
exposure to generate the photoexcited electrons and holes on the surface of the catalyst that would
allow the photoelectrocatalytic synthesis of hydrogen and oxygen, by reduction and oxidation of
water, respectively. Varieties of materials like oxides, sulfides, selenides, tellurides, nitrides, oxyni-
trides, oxy sulfides, and so on, of metals, are considered good materials for overall water-splitting
purposes.
The hydrogen produced would be the fuel to store the Sun’s energy, and it may be converted back
into electricity in a fuel cell, using the reverse reaction to water electrolysis. It is generally recog-
nized that one of the key fundamental restrictions in making the photoelectrochemical generation of
fuels a viable technology is the poor kinetics of the anode reaction (i.e., the water oxidation to oxy-
gen). Although sunlight-driven water splitting shows a potential route to sustainable hydrogen fuel
production, their widespread implementation has been hampered by the expense of photovoltaic and
photoelectrochemical apparatus and also by the stability of the catalytic material.
A search is on for a highly efficient and low-cost water-splitting cell combining a state-of-the-art
solution, in the form of the new materials in this field, such as bifunctional Earth-abundant catalyst
NiFe-layered double hydroxide in perovskite tandem solar cell (12.3%), gold electrode covered with
an iron-sulfur complex with doped layers of indium phosphide (InP) nanoparticles (60%), [Mo3S13]22
clusters, Fe-doped graphene oxide QDs (2.5%), bimetal oxynitride (5.2%), metallo-oxysulfides,
composite nanomaterials, etc. (Kibsgaard et al. 2014; Zeng and Zhang 2010; Luo et al. 2014). The
catalyst electrode exhibits high activity toward both the oxygen and hydrogen evolution reactions in
electrolyte on submerging in water and irradiating with light under the small electric current.

8.5.7 Hydrogen as a By-Product of Other Chemical Processes


Technological developments, economic conditions, and governmental policy set the trends that
aligned to create a significant market opportunity for waste-to-energy (WTE) plants, which utilize
the considerable amount of hydrogen as a by-product of industries, municipal solid waste (MSW),
and wastewater treatment plants for the production of electricity and heat. It can be generated as the
industrial by-product of diverse process industries (such as electrolysis in vinyl industries, the unsat-
uration process of hydrocarbons, distillation or dissociation of long-chain hydrocarbons in petro-
leum industries, chlorine industries, electrolysis of water, etc.) that can generate a sizable amount
of hydrogen by-product. In 2011, the United States developed 40,000 waste treatment facilities that
could be modified to generate hydrogen. A connection between energy source, type, technology,
and production method of hydrogen has been demonstrated (Table 8.1).
In addition to the above processes, there are a few more industrial processes that are used to
produce hydrogen as a by-product, as discussed below.
Hydrogen 153

TABLE 8.1
Hydrogen Production Using Diverse Forms of Technologies
and Energy Sources
Energy Source Energy Type Energy Technology Production Method
Renewable Power Wind Electrolysis
Hydro
Waves Photovoltaic
Sun
Fuel Biomass Reforming
Biochemical
Fossil Power Nuclear Electrolysis
Thermo
Fuel Natural Reforming
Oil
Gas Gasification
Coal

8.5.7.1  Catalytic Reforming of Oil Refinery


Dehydrogenation occurs during catalytic or thermal cracking of the petroleum. The residence time
of naphtha feedstock in traditional cracker furnaces is very low (less than milliseconds) to avoid the
formation of undesirable heavy hydrocarbon or coke formation, because the coke layer developed
inside the furnace reduces the heat transfer rate. The furnace is taken out of service, and decoking
is carried out with air and steam. Thus, the cracker furnace operates in a cyclic manner between
the cracking and decoking operation. Usually, steam is introduced with the feed to remove the coke
layer on the tube surface by converting coke into carbon monoxide and hydrogen by gas reaction
(Equation 8.25):

C + H 2O (steam ) → CO + H 2 (8.25)

8.5.7.2  Chlor-Alkali Plants


Chlor-alkali plants are based on the electrolysis of sodium chloride (brine water) for industrial
production of chlorine and sodium hydroxide (caustic soda). The H2 is created as a by-product in
manufacturing chemicals like chlorine and NaOH. A basic membrane cell was used in the elec-
trolysis of brine. At the anode, chloride (Cl−) ions were oxidized to chlorine. The ion-selective
membrane allows the counter-Na+ ion to freely flow across the membrane but prevents anions such
as hydroxide (OH−) and chloride from diffusing across. At the cathode, water is reduced to hydrox-
ide and hydrogen gas. The net process is the electrolysis of an aqueous solution of NaCl, which
produced industrially useful products sodium hydroxide (NaOH) (1.1 tons) and chlorine gas (1 ton)
(Equation  8.26). This overall reaction for every ton of chlorine produced during the electrolytic
process of brine, 28 kg hydrogen as a by-product is also generated. Four percent of total world-
wide hydrogen production is created by this process, using seawater (Figure 8.5 and Equation 8.26)
(Chouhan and Liu 2011):

Overall reaction: 2NaCl + 2H 2O → Cl2 + H 2 + 2NaOH (8.26)

8.5.7.3  Hydrocarbon Waste of High-Temperature Fuel Cells


High-temperature fuel cells based on molten carbonate (MCFC) or solid oxide (SOFC) technology
operate at sufficiently high temperatures and run directly on methane (Chouhan and Liu 2011).
This  is sometimes called “internal reforming.” Thus, MCFC and SOFC systems do not need a
154 Solar Energy Conversion and Storage

Ion-selective
membrane
Cl2 H2

28% NaCl Na+ Na+ H2O

Membrane

Cathode
Cl– H2 + OH–

Anode
Cl–
Cl2 H2O
OH–
26% NaCl NaOH

FIGURE 8.5  Ion-selective membrane cell for electrolysis of brine water for electrolysis of water. (Adapted
from Boettcher, S. W., E. L. Warren, M. C. Putnam, E. A. Santori, D. Turner- Evans, M. D. Kelzenberg, M. G.
Walter, J. R. McKone, B. S. Brunschwig, H. A. Atwater, and N. S. Lewis, J. Am. Chem. Soc., 133: 1216–1219,
2011. With permission.)

pure or relatively pure hydrogen stream as do proton exchange membrane (PEM) and phosphoric
acid (PAFC) systems, but these systems can run directly on natural gas, biogas, or landfill gas.
Furthermore, such systems can be designed to produce additional purified hydrogen as a by-prod-
uct, by feeding additional fuel and then purifying the hydrogen-rich “anode tail gas” from the fuel
cell into purified hydrogen.
Some cleanup of the methane stream may be required depending on the source of the methane.
However, projects such as the Ford Motors plant in Canada-FCE’s and Sierra Nevada Brewery in
Chico, California, have successfully demonstrated MCFC systems running on a blend of natural gas
and brewery wastewater treatment digester gas. These and other wastewater treatment and landfill
gas systems are attractive opportunities for renewably powered fuel cell systems.

8.5.7.4  Waste Biomass of Wine Industries


A typical winery in California can generate up to 10–12 million gallons of wastewater per year.
The Napa Wine Company of Oakville, California, produced hydrogen by microbial electrolysis. It
was the first out-of-laboratory renewable method for hydrogen production from wastewater using
a microbial electrolysis cell (MEC) system. The refrigerator-sized hydrogen generator takes win-
ery wastewater and specific bacteria along with a small amount of electrical energy to convert the
organic material into hydrogen. MEC is an electrolysis cell, where exo-electrogenic bacteria oxidize
biodegradable substrates and produce electrons and protons at the cathode and anode, respectively.
Hydrogen gas is produced at the cathode through a recombination of electrons with protons, assisted
by an additional voltage supplied by an external power source.
Experiments have determined that one group of bacteria turns unused sugar and unwanted
vinegar from improper fermentation into electricity at a working potential of around 0.3 V, which
is insufficient for the electrolysis of water. Theoretically, only an additional 0.11 V was needed
to produce hydrogen, but in practice, an additional 0.25 V is required due to overpotential at the
cathode (Logan and Regan 2006). This means that approximately half the energy needed for
electrolysis is possible with MECs, but overall system efficiency is to be explored. The complete
system includes maintaining MEC operating conditions, delivering feedstocks to the production
facility, replacing other expendable materials, and requiring gas separation and cleanup. One of
the biggest problems encountered with this project is maintenance of the effective concentration
of bacteria at a certain level that maintains the production rate. Another major issue related to this
technique is greater production of methane than hydrogen because hydrogen is being consumed
by “methanogenic” microbes before leaving the solution. This methane could then be “reformed”
into hydrogen by SMR, but the direct production would be far preferable from an overall energy
use standpoint.
Hydrogen 155

8.6  PHOTOCATALYTIC HYDROGEN GENERATION


It would be not overstated to call the twenty-first century the Age of Light (Fujishima and Tyrk
2012). Researchers and industrialists are looking for light-induced chemical and electrochemical
reactions that can be performed in an eco-benign way by reducing their energy demands. In this
context, annual sunlight exposure on the Earth’s surface contained an ample amount of energy—
that is, 3 × 1024 J, which is almost 104 times greater than the worldwide energy requirement or con-
sumption. Efforts are on to develop efficient materials that can convert solar energy into some useful
form of energy (electricity, chemical energy, etc.). This is one of the holy grails of future research
and technology development that can lead to the path of clean energy. The following criteria must
be satisfied, simultaneously, in designing or producing that kind of substance:

• Absorption of sunlight by some chemical substance that leads to the generation of elec-
trons and holes
• Effective separation of these photogenerated electron–hole pairs at little energetic loss
• Efficient charge transfer on the surface of the material that can promote oxidation and
reduction sites on the same surface
• Durability of material so that it should be stable enough at required conditions
• Apt energetics of the material that means their band edges must straddle between the redox
potentials of the substance (H2O, H2S, pollutant, dye, CO2 + H2O, NH3, etc.) to be broken
• Complementary metal-oxide semiconductor (CMOS) and biocompatibility
• Affordable cost of materials
• Nontoxic and easy-to-handle materials and sensitizers

Moreover, nature provides us a big clue about the enigmatic material in the form of the photosynthe-
sis process in living plants or in some bacteria. It is a remarkable example of solar energy conver-
sion into chemically stored energy. Lipids in chloroplast of green plants oxidized water into O2 and
reduced CO2 into useful organic chemicals in the presence of sunlight. Therefore, photosynthesis
fuels life on Earth by providing not only the carbohydrates (food) and oxygen (respiration) but also
the fossil fuels.
In looking to the potential of hydrogen fuel, there is a growing interest in storing chemical energy
or energy carriers as hydrogen. But fortunately or unfortunately, molecular hydrogen is not freely
available on Earth in convenient natural reservoirs. Most of the hydrogen on Earth is bonded to
oxygen in the form of water. Currently, elemental hydrogen is manufactured by the consumption of
a conventional steam reforming or decomposition of long-chain hydrocarbon present in fossil fuel.
In the former case, decomposition of water requires electrical or heat input, generated from some
primary energy source (fossil fuel, nuclear power, or renewable energy) at an industrial level. The
latter consumes the fossil resource and produces carbon dioxide, but often it requires no further
energy input beyond the fossil fuel.
We need first to invest energy to extract the energy that is contained in the hydrogen fuel. A
foremost method of hydrogen generation from fossil fuel, steam reforming, leads to carbon dioxide
emissions, the same as a car engine produces. One needs to think and act on the line of renewables
for hydrogen generation to substitute for fossil fuels, because we need to develop a method with less
dependence on fast depleting fossil fuel sources, stability in most regions of crude oil–possessing
countries, geopolitical independence, attributes to combat global warming. Therefore, producing
hydrogen in an eco-friendly manner is welcome. Moreover, nature provides us with the most abun-
dant resource of hydrogen in the form of water and free sunlight that can split water into hydrogen
and oxygen without releasing carbon content into the environment. This process can be named arti-
ficial photosynthesis. Worldwide, a very small portion of the hydrogen (4% in 2006) is produced by
electrolysis of water using electricity that consumes approximately 50 kWh of electricity per kilo-
gram of hydrogen production at the cost of the environmental degradation. Hence, the use of solar
156 Solar Energy Conversion and Storage

energy to generate hydrogen and oxygen out of water, is an attractive choice to meet the growing
demand for developing clean energy systems, without disturbing the eco-balance. Now the question
arises about the material, which can efficiently split the water. A number of series of semiconductors
have been tested on the criteria of cost, stability, and efficiency.
In the early 1970s, Fujishima and Honda (1972) used titanium dioxide (TiO2) as a first photocata-
lytic material for water splitting by achieving 1% efficiency. Since then, the titanium dioxide serves
as a model material for the mechanistic studies of photocatalysis, but it has a serious limitation as
a photocatalyst for water splitting. Its bandgap is too large (3.2 eV) for absorbing visible photons in
sunlight (46% Sun radiation); therefore, it is necessary to synthesize or modify the photocatalysts
workable with sunlight. Though the UV-based photocatalysts will perform better per photon effi-
ciency than visible light–based photocatalysts due to the higher photon energy, far more visible light
than UV light reaches the Earth’s surface. Thus, a less efficient photocatalyst that absorbs visible
light may ultimately be more useful than a more efficient photocatalyst absorbing solely light with
smaller wavelengths. In a single-step photocatalysis, the H2 evolution rate and STH energy conver-
sion efficiency are treated as a function of the bandgap energy of the photocatalysts with a 100%
quantum yield.
It is assumed that photons with wavelengths shorter than 1008  nm convert water to ½H2. If
the photocatalysts utilize visible light up to 600–700 nm with a quantum yield of 50%, the STH
efficiency reaches ~10%, which is similar to some of the commercial photovoltaic systems. But the
photocatalytic water-splitting devices are simpler than photovoltaic systems, because vacuum pro-
cessing is not required in the preparation of devices.
Copious materials of different categories have been studied, and research is still ongoing to
determine the visible light active proper material (efficiency more than 10%) for photocatalytic
water splitting. Usually, photocatalytic materials along with their highest oxidation states, contain-
ing metal cations with d0 and d10 electronic configurations, are suitable for overall water splitting.
Heterogeneous photocatalyst materials include metal oxides, nitrides, and sulfides, metal (oxy) sul-
fides, and metal (oxy) nitrides. These are rigorously tested on the criteria for better efficiency (Kudo
and Miseki 2009).

8.7  MECHANISMS OF PHOTOCATALYTIC WATER SPLITTING


The phenomena of water splitting followed four main steps (Figure 8.6):

• Sunlight with photonic energy more than 1008  nm is absorbed on the photocatalytic
surface.
• Charge carriers (electron and holes pairs) are created in conduction band (CB) and valance
band (VB), respectively.
• The grooves on the photocatalytic surface functioned as an anode (oxidation site) that
produced O2 using holes.

H2O

3 4 H2
e– e–
1 Co-catalyst
FIGURE 8.6  Four steps involved in photocatalytic water
Main catalyst
splitting for hydrogen production: (1) receiving photons of
2
energy more than 1008  nm: (2) Formation of photocarriers
(i.e., photoelectron and holes): (3) Separation of carriers by h+
movement of e− (reduction site co-catalyst) and hole toward O2 3
oxidation site (grooves on catalyst surface): (4) Reduction and H2O
oxidation of water at reduction and oxidation site, respectively.
Hydrogen 157

Engineered BG

0.00 eV

1.23 eV

Large BG Solid solution Narrow BG materials

FIGURE 8.7  Bandgap engineering in solid solution photocatalysts, where conduction band and valence
band are represented by white and gray boxes, respectively.

• Finally, electrons move toward the co-catalyst (anodic site) for reduction of water to pro-
duce hydrogen.

The reduction and oxidation reactions are the basic mechanisms of photocatalytic water splitting. The
CB level should be more negative than the hydrogen production level EH2/H2O = 0.00 eV for hydrogen
production, while the VB should be more positive than the water oxidation level EO2/H2O = 1.23 eV
for efficient oxygen production from water by photocatalysis (Figure 8.7). Theoretically, all types
of semiconductors that satisfy these requirements can be used as photocatalysts for hydrogen and
oxygen production through water splitting.
Although solar water splitting does not suffer from electricity storage problems, unlike the pho-
tovoltaics, it can provide two important feedstocks: hydrogen and oxygen. But their efficiency of
energy conversion has remained much lower than that of photovoltaics (Fujishima and Tyrk 2012).
The upgradation in efficiency based on the emergence of new materials with efficient solar energy
harvesting can be possible by lowering bandgap with the assistance of band position engineering,
nanostructuring, sensitization with dyes or quantum dots, and addition of proper dopant/alloy and
active co-catalyst to main photocatalyst (Bao 2014). The introduction of new systems like composite
materials, solid solutions, or plasmonic materials with their fine-tuned surface plasmon resonance
and supramolecular assemblies, as well as introduction of the active co-catalysts for hydrogen and
oxygen evolutions may enhance the efficiency of water splitting. Recently, one of the most important
advancements reported is in the form of the functionalized graphene oxides that can perform overall
water splitting without co-catalysts and a sacrificial reagent. Some of the representative examples
along with their water-splitting rates (WSR) and quantum yield (η) are as follows.

8.8  MODIFICATION OF PHOTOCATALYSTS


8.8.1  Photosensitization: Dyes/Quantum Dots
Sensitization is a widely used technique to utilize visible light for energy conversion. Some dyes
having redox property and visible light sensitivity can be used in solar cells as well as photocatalytic
systems (Choi et al. 1994; Maeda and Domen 2010). Under visible light illumination, the excited
dyes/quantum dots (QDs) can inject electrons to the CB of the photocatalyst to initiate the catalytic
reactions (Figure  8.8). Photoactivity of the whole system depends on the efficient absorption of
the visible light and efficient transfer of the electrons from excited dyes/QDs to the CB of TiO2.
Afterward, the CB electrons can be transferred to the co-catalyst (noble metal such as Pt) loaded
on the surface to reduce the water. Redox systems or sacrificial agents, such as the I3−/I− pair and
EDTA, can be added to rejuvenate dyes that sustain the reaction cycle.
The selection of QDs/dye-sensitized systems benefits the reaction by improving electron–hole sepa-
ration or suppression of charge recombination centers, thus improving the solar visible light–harvesting
power of the photocatalyst, and the selectivity or yield of a particular product. It has been shown that
the inorganic sensitizers, organic dyes, and coordination metal complexes can be used as the effective
158

(a) (b) Pristine ZnO NTs (c) (a) (b) CdSe


Light ZnO NTs@ CdSe(QDs)
S– on ZnO ZnO (QDs)
H 3O + 6 Annealed ZnO NTs@ CdSe(QDs) tube
Nanotube
CB S+
Pt 5
(c)
½ H2 aggregates
CdSe (QDs)
e-
S 4 (0.10)
H 2O (002) P a (0.11)
3 c
lane

(0.02)
0.26 nm

(mA cm–2)
Tio2 ½ O2 + 2H+ 2 (d)101
Light
1 Off m
5.0nm 6n

Photo current density


0.12 nm
VB 0 (e) (f)
0 100 200 300 400 500 600
Time (s)

FIGURE 8.8  (a) Dye sensitization of the photocatalyst for overall water splitting. (b) Efficient and sustainable QD CdSe sensitized ZnO nanotubes used for water
splitting. (c) High-resolution transmission electron microscopy (HRTEM) image, showing the lattice fringes of CdSe(QDs) loaded on ZnO NTs; transmission electron
microscopy (TEM) image of aggregates of CdSe QDs loaded on 002 plane of ZnO nanotubes; SAED pattern of pristine ZnO NTs and HRTEM lattice image of single
CdSe QD. (Adapted from Chouhan, N., C. L. Yeh, S. F. Hu, R. S. Liu, W. S. Chang, and K. H. Chen, Chem. Commun., 2011, 47, 3493–3495. With permission.)
Solar Energy Conversion and Storage
Hydrogen 159

photosensitizers (Gurunathan et al. 1997; Sauve et al. 2000; Jing and Guo 2007). QD sensitization
with their significant qualities like quantum confinement, efficient charge transfer, and separation are
attracting the attention of researchers. Photosensitization by dyes is common in solar cells due to their
prominent visible light absorption properties, ease of change in their structures, and low cost.

8.8.2 Reducing Bandgaps Through Doping


Noble metal, transition metals, or nonmetals (B, C, N, F, P, S) alloying/doping to the main photo-
catalytic system is also a popular mode of reducing bandgaps of the photocatalysts by adjusting
Fermi levels (Ni et al. 2007; Zhang et al. 2013) (Figure 8.9).

8.8.3 Engineered Solid Solutions


Solid solutions are usually produced by the combination of large-bandgap materials with low-
bandgap solids at high temperature (Figure 8.7). These solutions contain four or more elements. The
(Ga1−xZn x)(N1−xOx) and Ta-based semiconductors are well-known examples of such semiconductor
alloys. UV-based photocatalyst NaTaO3:La (WSR = 9.7 mmol h−1 and η = 56%) (Kato et al. 2003)
and K3Ta3B2O12 (WSR = 1.21 mmol h−1, and η = 6.5%) without co-catalyst assistance and use of
sacrificial reagents (Kurihara et al. 2006), yields pretty high efficiency for water splitting. Visible
light active Ga0.82Zn0.18N0.82O0.18 (WSR = 0.4 mmol h−1 and η = 5.9%) (Maeda et al. 2008), loaded
with co-catalyst Rh2–yCryO3 and without utilizing the sacrificial reagents, has the highest quantum
yield along with a water-splitting rate in the presence of visible light. These catalysts can be tuned
for high efficiency by calcination that reduced the amount of surface defects, which can act as
recombination sites.

8.8.4 Macromolecular Systems for Water Splitting


Thermodynamics of the solar water splitting involves an energetically uphill chemical reaction that
required 1.23 V. Water reduction is energetically favorable via a multielectron pathway (1.23 V vs.
NHE) compared to via a single-electron pathway (5 V vs. NHE) that makes the design of a complete
water-splitting system quite complicated. Therefore, an in-depth understanding of either oxidation or
reduction by replacing the other half of the reaction with a sacrificial electron acceptor (EA) and an
electron donor (ED), is required. Many macromolecular systems are emerging as the photocatalytic
active substances that include organized zeolites-based photoredox systems. A synthetic strategy for
the construction of zeolite-entrapped organized molecular assemblies has to be developed. This can
be achieved by interacting adjacent cage dyads composed of two polypyridine complexes of Ru(II)

(a) (b) (c)


1.0
Ti0.INb 650°C
Photocurrent density

0.8

0.6
(mA cm–2)
F (R)

0.4 4 C Coating/H-TiO2
Current density

TiNT 650°C
(mA cm–2)

0.2
2
0.0

–0.2 0
–1000–800 –600 –400 –200 0 200 400 600 0 250 500 750
λ (nm) Potential (mV) Time (s)

FIGURE 8.9  (a) Modification of UV photocatalysts by cation/anion doping. (b) Nb-doped TiO2 nanotube
versus Ti–Nb alloys, showing strongly increased and stable photoelectrochemical water splitting. (c) Enhanced
water-splitting performance of TiO2 nanotube arrays coated with an ultrathin nitrogen-doped carbon film.
(Adapted from Das, C., P. Roy, M. Yang, H. Jha, and P. Schmuki, Nanoscale, 3, 3094–3096, 2011, and Tong,
X., P. Yang, Y. Wang, Y. Qin, and X. Guo, Nanoscale, 6, 6692–6700, 2014. With permission.)
160 Solar Energy Conversion and Storage

or from heterosupramolecular chemistry, metal clusters, dendrimers, and so on (Sykora et al. 1998;
Arachchige et al. 2009). Among them, the most common photocatalysts with normal structure and
diverse function solar hydrogen–producing capacity, are the systems that include Ru, Pt, Ru–Rh, and
Rh–Rh supramolecular architecture.
In this context, supramolecular chemistry provides an ideal solution to study, design, and
develop a solar hydrogen photocatalyst for overall water splitting. Overall photocatalytic activi-
ties of the supramolecules involve some individual component, which performs a specific act of
either oxidation or reduction. Consequently, the assembly of supramolecules can be varied in
a fine-tunable fashion, along with maintenance of the overall water-splitting process. At least
two electrons are collected at the central site of the supramolecule after optical excitation by
means of a photoinitiated electron collection (PEC) process. For an efficient PEC, supramol-
ecules can be designed to have a TL/ED-LA-BL-EC-BL-LA-ED/TL structural motif, where the
term TL is a terminal ligand (e.g., two bpy, two dpp, two dpb, two phen, or tpy and Cl ligands;
bpy = 2,2′-bipyridine, phen = 1,10-phenanthroline, tpy = 2,2′:6′,2″-terpyridine, dpp = 2,3-bis(2-
pyridyl)pyrazine, dpb = 2,3-bis(2-pyridyl) benzoquinoxaline), LA is light absorber, BL is bridging
ligand (dpp, whereas the Ir(III) system uses the dpb ligand), and EC is electron collector (M′X2;
M′ = Rh(III) or Ir(III) and X = Cl or Br).
The complexes possess intense MLCT transitions in the visible region of the spectrum, making
them efficient LAs. Light absorption is dictated by the TL-LA-BL subunit. The presence of two LAs
connected to a single EC through bridging ligands allows electron collection with multiple photoex-
citation steps and intense metal ligand charge-transfer transitions in the visible region of the spec-
trum, making them efficient LAs. Light absorption is dictated by the TL-LA-BL subunit. Thus, the
supramolecular motifs have great potential to exploit the light radiation for breaking H2O for hydro-
gen generation. The first functioning PEC material was [(bpy)2Ru(dpb)2IrCl2][PF6]5, which collects
two electrons on the dpb(π*) orbital upon sequential excitation and electron transfer events. The
three most active photocatalysts, [{(bpy)2Ru(dpp)}2RhCl2](PF6)5, [{(bpy)2Ru(dpp)}2RhBr2](PF6)5,
and [{(phen)2Ru(dpp)}2RhCl2] (PF6)5 were used for the higher hydrogen yield using the DMA elec-
tron donor (Balzani et al. 1997). The results of the screening of these new photocatalysts reveal that
the designs are functional for supramolecular complexes as photocatalysts in general or complexes
that can be predictable on the basis of their basic chemical properties. For systems that do function,
optimization of the reaction conditions, by way of component modi-cation, requires identifying new
electron donors and their optimized experimental conditions.
Dedicated efforts of the researchers in this direction are underway to gear the better PEC in the
supramolecular architecture for achieving higher hydrogen generation efficiency. The first rhodium-
centered trimetallic supramolecular [{(bpy)2Ru(dpp)}2RhCl2]5+ was reported that collects multiple
electrons and is capable of photocatalyzing hydrogen production from water. This supramolecular
system incorporated two [(bpy)2Ru(dpp)]2+ LAs and a single Rh EC, which allowed for the multiple
electron collection necessary for the multielectron reduction of the substrate. The monodentate
Rh-bound ligands can be lost after photoreduction of the substrate, which provides a site for sub-
strate binding for photoreactivity to the supramolecular complex.
The redox properties of [(bpy)2Ru(dpp)2RhCl2]5+ demonstrated that the Ru LA are largely elec-
tronically isolated important PEC. Its multielectron photochemical component was modified by
changing the terminal ligand (TL), TA metal, and Rh-bound halides. Subsequently, it afforded a
series of additional photoinitiated ECs having a LA-BL-Rh2X-BL-LA structure motif (LA = Ru″
or Os″ polyazine X = Cl−, Br−, I−, BL-dpp), and these cells are capable of catalyzing solar hydro-
gen production. Studies have established [(bpy)2Ru(dpp)2 RhCl2]5+ and [[(phen)2Ru (dpp)]2RhX2]5+
X = Cl−, Br−, as photochemical molecular devices for PEC hydrogen production photocatalysis with
a hydrogen yield ϕ = 0.01 (Dincer and Zamfirescu 2012). Ozawa et al. (2006, 2007) investigated a
[PtII(bpy)2Cl2] component linked to a ruthenium LA through an amide linkage to ruthenium capable
of photochemically producing hydrogen from water (ϕ ≈ 0.01 and 5 turnovers in 10 h) in the pres-
ence of ethylenediaminetetraacetic acid. Rau et al. (2006) reported a Ru-Pd bimetallic system that
Hydrogen 161

photochemically produces hydrogen in the presence of an electron donor, TEA, affording 56 turn-
overs in ∼30 h.
Recent studies on platinum (Du et al. 2008) and palladium (Lei et al. 2008) systems suggested
that hydrogen production can be attributed to the decomplexation of the reactive metal to form
colloidal metal, which can function as the hydrogen generation site. Elvington and Brewer (2006)
reported the photoinitiated electron collection by a metal center in [{(bpy)2Ru(dpp)}2RhCl2](PF6)5.
When excited with visible light, this complex is reduced, converting Rh(III) to Rh(I), which under-
goes chloride loss to form [{(bpy)2Ru(dpp)}2RhI]5+ (Elvington et al. 2007). This complex photo-
chemically reduces water to hydrogen using visible light excitation with ϕ ≈ 0.01 in the presence
of a sacrificial electron donor DMA. The ability of the system to undergo photoinitiated electron
collection on the rhodium center with the molecular architecture remaining intact is unprecedented
and allows its use in multielectron photochemistry.
The design considered for mixed-metal supramolecular complexes capable of photoinitiated
electron collection provides an insight into general molecular architectures for photocatalysts for
solar energy–driven hydrogen generation from water. Some of the polymetallic complexes that
couple ruthenium or osmium polyazine LAs to rhodium or iridium metal centers through BLs have
been reported with many interesting light-activated DNA photocleavage agents (Elvington et al.
2007; Arachchige et al. 2008). Existence of the supramolecular catalyst in the photocatalytic world
is prominent, but it is a major challenge to maintain photoelectrons for a sufficiently long duration
at the supramolecular level.

8.8.5  Plasmonic Nanostructures with Surface Plasmon Resonance


Nanoparticles of pure metals show active absorption in the visible region because of the existence
of the unique feature known as surface plasmon resonance (SPR). Plasmonic photocatalyst has two
prime parts (1) nanoparticles of noble metals with LSPR and (2) support/carrier (semiconductor/
polar material) with Schottky junction. Nanoparticles of metals contribute strong absorption of vis-
ible light and excitation of active charge carriers, whereas the latter induces the charge separation
and transfer. Free electrons of the noble metals will integrate with the photon energy that produces
subwave and conducting electrons in oscillating mode and offers localized surface plasmonic reso-
nance (Ritchie 1957).
Resonance frequency of the plasmonic material can be tuned by varying the nanoparticle size,
shape, material, and proximity to other nanoparticles. Due to LSPR, NPs of noble metals can act
as the thermal redox reaction-active centers on catalyst that can trap, scatter, and concentrate light.
Furthermore, integration of the plasmonic metals with the semiconductor includes plasmon reso-
nance energy transfer (PRET), photon scattering effect, hot electron transfer, plasmon-induced heat-
ing, reflection reduction, reduced e/h diffusion length, enhanced local electric effect and molecular
polarization effect, quantum tunneling effect, and high catalytic effect, at the metal/semiconductor
interface (Barnes et al. 2003; Maier and Atwater 2005; Maier 2007; Lal et al. 2007). The typical
examples of exhibiting these properties are plasmonic silver or gold nano particle–assisted oxide
semiconductor N-TiO2 and CdS@SiO2//Au@SiO2 (Figures 8.10 and 8.11).
In typical PEC water splitting with plasmonic metal integrated metal oxide, hot electron of noble
metal is injected to conduction bands of metal oxides, which increases the power absorbed in this
region and enhances the efficiency of the device. Few of the plasmonic photocatalytic systems
for enhancement of the renewable hydrogen evolution through photocatalytic water splitting are Au/
graphene/TiO2 (12 mmol/2.4 mmol h−1) (Singh et al. 2014), Au/ZnO nanorods (11.2 and 4.4 μmol h−1)
(Chen et al. 2012), CdS@SiO2//Au@SiO2 (10 mmol h−1 in 300 min exposure) (Torimoto et al. 2011),
and many more. By integrating plasmonic materials with hydrogen and oxygen co-catalyst on the
main catalyst, overall water splitting will be further enhanced. Because the scattering of photons
by plasmonic nanoparticles promotes PRET phenomena that allow efficient electron transfer and
lead selective formation and separation of the electron–hole pairs in the near-surface region of the
162 Solar Energy Conversion and Storage

(a) (b) Optical overlap of building blocks (c) (d)


Blank
0
Ag
Au
N-TiO2 absorbance Ag
TiO2
Ag/TiO2 Ag cubes

Extinction [a.u.]

Extinction [a.u.]
N-TiO2
Ag SPR Au spheres
ln (C/C0)

Au/N-TiO2
–0.1
Au SPR 1 µm

–0.2 Au
N-TiO2

Ag/N-Tio2 Minimal overlap


Large overlap TiO2
–0.3
0 10 20 30 300 400 500 600 700 800 200 nm 300 400 500 600 700 800
Time [min] Wavelength [nm] Wavelength [nm]

FIGURE 8.10  (a) Aqueous methylene blue decomposition over different photocatalysts: Ag cubes only, Au
spheres only, P25 TiO2 only, N-TiO2 only, Ag/TiO2 composite, Au/N-TiO2 composite, and Ag/N-TiO2.
(b)  Overlap between the illumination source spectrum, semiconductor absorbance spectrum, and metal
nanoparticle surface plasmon resonance spectrum. (c) Scanning electron micrograph (SEM) of the Ag nano-
cube (edge length = 118, 25 nm) and spherical gold particles (diameter 25.4, 4.5 nm). (d) Diffuse reflectance
UV-visible spectra of TiO2, N-TiO2, Ag nanocubes, Au spheres. TiO2 and N-TiO2 source. (Adapted from
Ingram, D. B., P. Christopher, J. L. Bauer, and S. Linic, ACS Catal., 1, 1441–1447, 2011. With permission.)

(a) SiO2 CdS


(d) 12
MPTS-CdS
Au Au
Amount of H2 (μmol)

Hydrolysis
Au@SiO2 8
CdS@SiO2//Au@SiO2

(b) (c) 4

0
0 100 200 300
Irradiation time (min)
50 nm 50 nm

FIGURE 8.11  (a) Immobilization of CdS@SiO2 on Au@SiO2 particles. (b) TEM images of Au@SiO2 pre-
pared with Au particles of 19 nm in diameter as a core. (c) CdS@SiO2-deposited Au@SiO2 particles CdS@
SiO2//Au@SiO2. (d) Time courses of hydrogen evolution by photocatalysts of CdS@SiO2 (solid circles) and
CdS@SiO2//Au@SiO2 particles with SiO2 layer thicknesses of 17 nm (open circles) and 2.8 nm (open squares)
on Au cores. The experiments were performed by irradiation light from a 300  W Xe lamp (λ  >  350  nm).
(Adapted from Torimoto, T., H. Horibe, T. Kameyama, K. Okazaki, S. Ikeda, M. Matsumura, A. Ishikawa,
and H. Ishihara, J. Phys. Chem. Lett., 2, 2057–2062, 2011. With permission.)

semiconductor. In totality, one can say that water splitting via SPR is a complex multifaceted pro-
cess, and enhancement in its efficiency requires the thoughtful attention to various parameters such
as morphology, stability, structural defects, bandgap and band-edge position of the metal oxides, and
size, morphology, plasmonic particle areal density, and location of plasmonic metal nanoparticles.

8.8.6 Nanostructuring of the Photocatalyst


Nanoparticles continue to exhibit much higher activity than their bulk counterparts due to the high
surface-to-volume ratio. These two-dimensional nanomaterials have the potential to be excellent
catalysts, as they can harvest solar energy and generate photoelectrons and holes, along with provide
paths for the separation and diffusion of photoexcited carriers. Liao et al. (2014) studied the cobalt
(II) oxide (CoO) nanoparticles that can carry out overall water splitting with a solar-to-hydrogen
Hydrogen 163

efficiency of around 5%. Bismuth-based systems have also been demonstrated to have an efficiency
of 5% with the advantage of a simple and cheap catalyst (Abdi et al. 2013). Furthermore, the p-type
conductivity results in the formation of an accumulation layer at the graphene (GO)/water interface,
which is favorable for water reduction to hydrogen. Modification of a graphene sheet to exhibit both
p- and n-type conductivities may produce a photocatalytic medium, which is effective for overall
water splitting into H2 and O2. Here, effective exciton separation and charge transfer are essential
factors responsible for overall water splitting. Reducing the size of the GO sheets may lower the
recombination probability of the photogenerated charges (Yeh et al. 2014).

8.8.7  Composite Photocatalyst


Oxide photocatalysts are the most commonly used photocatalysts in various applications of opto-
electronic materials due to their potential thermal and chemical stability, low cost, and environmen-
tally friendly aspects, but these are UV light–active materials due to their wide bandgap. Therefore,
some suitable nonoxide/oxide with a proper band-positioned photocatalyst is used to couple with
them to make them visible light active. The whole assembly of these systems is known as a com-
posite material. These were found to be suitable for water splitting because of their excellent pho-
tocorrosion resistance, which is ascribed to a considerable decrease of surface defect density on the
photocatalytic surface and the reduction of holes. A few of their examples are CdS-TiO2, CdS-ZnO,
CdS-AgI (Yamada et al. 2005), ZnO/Ag/CdS (η = 3.5 mL h−1) (Zhang et al. 2014), ZnO/Ag/AgCl
(Pirhashemi and Yangjeh 2014), and so on. The electron transfer mechanism for water splitting has
been demonstrated in Figure 8.12.
Material with more than 10% efficiency would be a major breakthrough for hydrogen produc-
tion by using solar water splitting (Tachibana et al. 2012). Recent progress toward developing arti-
ficial photosynthetic devices, together with their analogies to biological photosynthesis, including
technologies that focus on the development of visible light active heterostructures still require an
understanding of the underlying interfacial carrier dynamics. Of course, it has definitely proposed a
vision for a futuristic development of the novel efficient materials along with the improved designs
of the rational device, which will be based on theory and experimental simulations.

8.8.8  Photoelectrochemical Water Splitting


Electricity produced by photovoltaic systems offers one of the cleanest ways to produce hydrogen.
We can use photonic energy from sunlight to subsidize its electricity demand. These reactions are
known as photoelectrochemical (PEC) reactions. When water is broken into hydrogen and oxygen
by electrolysis under light exposure in a PEC device, the process is named artificial photosynthesis.
These PEC devices are made of three electrodes: photoanodes; photocathodes, mainly platinum

(a) H2 (b) Potential/eV


Phosphate Acceptor
e– Reduction
H2O Acceptor –1
H2/H+ e– CB’ 0 ~550 nm
Visible CB ~620 nm
light +1
+2
Ox/red VB
O2/H2O h+ Donor VB’ +3
Oxidation+
H2O h+ Donor Fe2O3 TiO 2
O2 CB: Conduction band bottom of Fe2O3 CB: Conduction band bottom of TiO2
VB: Valence band top of Fe2O3 VB: Valence band top of TiO2

FIGURE 8.12  (a) Water splitting in a composite system. (b) Improved photoactivity of TiO2–Fe2O3 nanocom-
posites for visible light water splitting after phosphate bridging and its mechanism. (Adapted from Luan, P.,
M. Xie, X. Fu, Y. Qu, X. Sun, and L. Jing, Phys. Chem. Chem. Phys., 17, 5043–5050, 2015. With permission.)
164 Solar Energy Conversion and Storage

Bias O2
H2

Proton permeable membrane


Reference electrode
Photocathode
Inlet

Quartz window
Water jacket 4H+ 4H++O2

source
Anode

Light
H+ H+ 2H2O
Outlet H2

FIGURE 8.13  Photoelectrochemical cell utilized for water splitting.

wires; and reference electrodes (calomel or Ag/AgCl electrode) dipped in aqueous electrolyte
(Figure 8.13). Photoelectrochemical solar devices must accomplish several tasks during water split-
ting. The steps involved in water splitting are

1. The photocatalyst absorbs short-wavelength, high-energy photons on a catalytic surface


and transmits longer-wavelength photons.
2. The mobile photocharge carriers (i.e., photoelectrons and holes) are generated.
3. The generated charge carriers are transported toward catalytic reduction and oxidation
sites.
4. The H2 and O2 are produced at the cathode and anode, respectively.

Electrochemical mass transport can be categorized according to three driving forces: convec-
tion, migration, and diffusion. Mechanical movement of the bulk solution generates convective
transport, which is suppressed by structured photoelectrodes. Migration refers to the motion of
charged species under the influence of electric fields, the operation in either highly acidic or highly
alkaline media, and the excess of the supporting electrolyte, which effectively suppresses ion migra-
tion forces. Ultimately, the diffusion is likely to dominate mass transport in the vicinity of struc-
tured photoelectrodes, as used for water-splitting purposes.

8.9  ELECTROCHEMICAL ASPECT


When a nonilluminated semiconductor electrode is in contact with the solution of a redox couple
electrolyte, then equilibrium taking place at the interface and the Fermi level of the semiconductor
can be adjusted with the same redox couple electrolyte. During this process, two layers are formed.
The first is the space charge layer/depletion layer (1–103 mm, depending upon the charge density
and dielectric constant) formed inside the semiconductor of the thin region close to its surface.
Mobile charge carriers have diffused away, or have been forced away by an electric field in the space
charge layer. The only elements left in the depletion region are ionized donor or acceptor impurities.
The second layer is the interfacial layer between the electrode surface and the electrolyte solution.
Bending (upward for n-type) or (downward for p-type) changing the direction of electronic bands
(valence and conduction bands) occurs at the space charge layer. The size of the bending depends
on the relationship between the Fermi level of the semiconductor and the redox couple solution, as
depicted by Figure 8.14.
If the semiconductor is illuminated by the light having photonic energy greater than the bandgap
of the semiconductor, then the electron–hole pair is generated and separated in the space charge
Hydrogen 165

(a) (b)
Equilibrium, dark;
Short circuit, illuminated Electrolyte Open circuit, illuminated Electrolyte
Electrode Space Electrolyte Space
Fermi level

Interfacial Layer
Eoc

Interfacial layer
Fermi level EF=0, of
Charge EF=0, Eg Charge
of counter- counterelectrode
Eg
electrode

Region Region

FIGURE 8.14  Energy level of generic n-type semiconductor in a representative PEC: (a) equilibrium, dark
and short circuit, illuminated condition; (b) open circuit, illuminated condition.

layer. In an n-type semiconductor, the electric field exists across the space charge layer, it drives
holes to move toward the interfacial layer (electrolyte/solution), and the electron moves to the inte-
rior of the electrode and from there to the external circuit. In an open circuit, electrons flow toward
the bulk semiconductor and begin to accumulate, which will raise the Fermi level. This level will
continue to increase depending on the light intensity and recombination rate, until a steady state is
reached. There is no net change in the number of oxidized and reduced forms of the redox couple
because the same number of oxidized forms is reduced at the semiconductor surface as the number
of reduced forms produced by the oxidized redox couple.
At illuminated semiconductor–liquid junctions, the open-circuit photovoltage, Voc represents a
fundamental figure of the merit for photoelectrochemical water splitting (Equation 8.27). As the
flow of current is a kinetic phenomenon, Voc is a kinetic parameter, not a thermodynamic parameter
(Boettcher et al. 2010):

K BT ⎛ J ph N D:A Lmin ⎞
Voc = n ln ⎜ 2 ⎟ (8.27)
q ⎝ qDmin ni ⎠

where:
n is the diode ideality factor
Jph is the short-circuit photocurrent density (photocurrent per unit area) under illumination
ND (NA) is the donor (acceptor) density
L min is the minority carrier diffusion length
Dmin is the minority carrier diffusion coefficient
ni is the intrinsic carrier density

The term saturation current density Js, is equal to –qDminni2/ND:A L min, as demonstrated by
Equation 8.28 (Boettcher et al. 2011):

K BT ⎛ J ph ⎞
Voc = n ln ⎜ ⎟ (8.28)
q ⎝ Js ⎠

The dimensionless figure γ is introduced in Equation 8.28 for structured materials used in bulk
photoelectrochemical water splitting, to reconcile the incongruity between the two areas (i.e., Jph
and Js) (Equation 8.29) (Stempel et al. 2008):

Geometric surface area


γ= (8.29)
Projected surface area
166 Solar Energy Conversion and Storage

Incorporation of term γ into Equation 8.29 enables a straightforward comparison of the currents and
current densities (Equation 8.30):

K BT ⎛ J ph ⎞
Voc = n ln ⎜ ⎟ (8.30)
q ⎝ γJ s ⎠

Equation 8.30 was introduced for the small contact area heterojunctions, optical concentrators, or
a nanoemitter-styled photoelectrochemical cell (Yamada et al. 2005; Zhang et al. 2014; Pirhashemi
and Yangjeh 2014).
Ayers (1983) of Energy Conversion Devices (United States), has demonstrated and patented the
first multijunction high-efficiency photoelectrochemical system for direct splitting of water. This
group demonstrated wireless solar water splitting with low-cost thin-film amorphous silicon mul-
tijunction sheets with metal substrate at back and sandwiched Nafion membrane (provided a path
for ion transport), merged in water. The H2 was evolved on the front catalyst–decorated amor-
phous silicon surface, while oxygen was evolved off the back of the metal substrate. Research
continues toward developing high-efficiency PEC technology at academic and industrial levels.
Few ambassador PEC electrode materials are mentioned here to cover the whole range of the
photocatalytic semiconductors. In particular, doped and undoped oxide photocatalyst ZnO (Ahn
et al. 2007, 2008; Yang et al. 2009) are used for water splitting because of their unmatched piezo-
electric properties. Anodically fabricated TiO2 nanotube arrays have proved to be robust and
cost-effective functional materials. These have been widely investigated in many applications,
especially those related to energy conversion, such as photoelectrochemical water splitting and
solar cells (Murphy 2008; Grimes and Mor 2009; Allam and Grimes 2009; Hamedani et al. 2011;
Sharmoukh and Allam 2012).
Simultaneously, the compositional doping of TiO2 with different elements was considered
as an approach for bandgap engineering (Sharmoukh and Allam 2012). The (Ga1–xZn x)(N1–xOx)
modified with Rh2–yCryO3 holds the highest AQY for overall water splitting using a single photo-
catalyst under visible light (5.1% at 410 nm) (Maeda and Domen 2010). However, solid solution
(Ga1–xZn x)(N1–xOx) did not exhibit good activity beyond 440  nm for water-splitting devices that
can typically harvest visible light and have a low solar-to-hydrogen efficiency of around 0.1%.
Co-based photocatalysts are emerging materials for hydrogen evolution with a solar-to-hydrogen
efficiency of 5%. They add volume to the developing potential of Co-based photocatalysts (Liao
et al. 2014). Additionally, some chalcogenides containing CuI ions, such as CuGaSe2, Cu2ZnSnS4,
and Cu(Ga,In)(S,Se)2, function as p-type semiconductors and can also be employed as photovoltaic
cells (Ma et al. 2011; Yokoyama 2010; Moriya et al. 2013). IrO2-loaded LaTiO2N electrodes have
been used with a Ta contact layer between the LaTiO2N particles and the Ti conductor layer in the
electrolyte of aqueous 1 M Na2SO4 solution with pH = 13.5 (adjusted by the addition of NaOH)
(Minegishi et al. 2013).
CoOx was also found to be effective for nitride and oxysulfide photocatalysts such as Ta3N5
(Ma et al. 2013) and Sm2Ti2S2O5 (Ma et al. 2012). CdS is an n-type semiconductor and can form a
good p-n heterojunction with CuGaSe2. The photoelectrode surface was modified with Pt to cata-
lyze the hydrogen evolution reaction. The band bending initiated in CdS on CuGaSe2 is the key to
enhance the photocurrent (Figure 8.15a) (Moriya et al. 2013). The highest AQY of 6.3% at 420 nm
was achieved with a Z-scheme system based on Pt-loaded ZrO2/TaON as a hydrogen evolution
photocatalyst, Pt-loaded WO3 as an oxygen evolution photocatalyst, and IO3−/I− as a redox couple
(Figure 8.15b) (Maeda et al. 2010), which has shown almost double activity.
Scientists are also interested in graphene species, a promising material for synthesizing metal-
free, cost-effective, and environmentally friendly catalysts for overall water splitting under solar
illumination. Although the water-splitting activity of the nitrogen-doped graphene oxide quantum
Hydrogen 167

(a) (b)
Pt
Light OFF Pt/CuGaSe2
Pt λ = 420.5 nm
e– e–
λ = 420.5 nm e– H2
Pt/CdS/CuGaSe2
Current density

e–
l– H+
H 2O
5 mAcm–2
lO3– e–
Light ON O2 e– h+
h+
5 mVs–1 AQY:~6.3%

–0.1 0.0 0.1 0.2 0.3 0.4 0.5 0.6 0.7 0.8 Pt/WO3 Pt/ZrO2TaON
(O2 evolution system) (H2 evolution system)
Potential/V vs. RHE

FIGURE 8.15  Photoelectrochemical: (a) current density versus voltage curve for Pt/CuGaSe2 and Pt/CdS/
CuGaSe2. (Adapted from Moriya, M., T. Minegishi, H. Kumagai, M. Katayama, J. Kubota, and K. Domen,
J. Am. Chem. Soc., 135, 3733–3735. With permission.) (b) Z-scheme based on Pt-loaded ZrO2/TaON as a hydro-
gen evolution photocatalyst, Pt-loaded WO3 as an oxygen evolution photocatalyst, and IO3−/I− as a redox couple.
(Adapted from Maeda, K., M. Higashi, D. Lu, R. Abe, and K. Domen, J. Am. Chem. Soc., 132, 5858–5868. With
permission.)

dots was reported to be almost half that of the Rh2–yCryO3/GaN:ZnO, it has a lot of possibilities for
the future (Maeda et al 2010; Yeh et al. 2014). The photocurrent densities of the gold nanorod–TiO2/
platinum electrode correspond to the hydrogen and oxygen production under the full sunlight irra-
diation. This system produced 5 × 1013 H2 molecules cm−2 s−1 under 1 sun illumination (AM 1.5 G
and 100  mW  cm−2), with unprecedented long-term operational stability. This is high among the
reported plasmonic systems in terms of solar-to-hydrogen efficiency (≈0.1%) but low enough for
practical use (Mubeen et al. 2013).

8.10  HYDROGEN AS A KEY SOLUTION


In looking to the recent developments in hydrogen production, storage, and potential as a fuel and
energy carrier, one can easily say that hydrogen will be a key solution to the current energy chal-
lenges faced by the different facets of economy/society. Hydrogen economy as a primary energy
source and feedstock produced from sources other than coal, oil, and natural gas, would cut down
the release of the greenhouse gases into the atmosphere, which is the primary feature associated
with the combustion of fossil energy resources. The major benefit of the hydrogen economy would
be that energy generation and its use could be decoupled. Therefore, the primary energy source
would not have to travel with the vehicle, as currently required for hydrocarbon fuels. Instead of
tailpipe-dispersed energy emissions, the energy could be generated from point sources, which are
otherwise unfeasible for mobile applications.
Centralized energy plants promise higher hydrogen production efficiency, but suffer from dif-
ficulties in high volume and long-range hydrogen transportation (due to factors such as hydrogen
damage and the ease of hydrogen diffusion through solid materials), which makes electrical energy
distribution attractive within the hydrogen economy. Under the distributed energy generation
schemes, the small-scale renewable energy sources could be associated with hydrogen stations. The
proper balance between hydrogen distribution and long-distance electrical distribution is one of the
primary questions raised about the hydrogen economy. Again, the dilemma of production sources
and transportation of hydrogen can now be overcome using on-site (home, business, or fuel station)
generation of hydrogen from off-grid renewable sources.
Distributed electrolysis would bypass the problems of distributing hydrogen by distribut-
ing electricity instead of hydrogen. Generated electricity can be transported to the end user by
using an existing electrical network or on-site electrolyzer located at filling stations. Hydrogen
168 Solar Energy Conversion and Storage

has been explored as an excellent fuel for passenger vehicles, where it can be used in fuel cells
to power electric motors or burned in internal combustion engines (ICEs). In both ways, it is
an environmentally friendly fuel that has the potential to dramatically reduce our dependence
on imported oil. The alternative fuel hydrogen falls into a category of nonconventional or
advanced fuels.
The combination of the fuel cell and electric motor is two to three times more efficient than an
ICE. However, the high capital cost of fuel cells is one of the major obstacles in its development that
has to be overcome in the future. Until then, the fuel cells are technically but not economically more
efficient than an ICE (Tester et al. 2005). Other technical obstacles include hydrogen storage issues
and the purity requirement of hydrogen used in fuel cells. With current technology, an operating
fuel cell requires the purity of hydrogen to be as high as 99.999%. Hydrogen engine conversion tech-
nology is more economical than fuel cells. Hydrogen produces no air pollutants or greenhouse gases
when used in fuel cells; however, it produces only nitrogen oxides (NOx), when burned in ICEs.
There are several significant challenges (price, hydrogen storage issues, and the purity of hydrogen
for fuel cells) and until solutions are available for these problems, it cannot be widely used. In the
event that fuel cells become price competitive with ICEs and turbines, large gas-fired power plants
could adopt this technology. Hydrogen can also be utilized as the raw material in many chemical
industries, cooking gas, fuel for house warming, fuel cells, power production, high-temperature
welding, and vehicles. Therefore, hydrogen is a potential candidate for providing a clean and green
energy solution to our present and future energy needs.
To sum up the story on hydrogen energy, it is worthy to summarize the benefits and drawbacks
of this alternative source of energy as a fuel and energy carrier. Furthermore, a hydrogen economy
should be duly established and its importance in present and futuristic points of view is to be stressed.
Hydrogen is popular for its high energy density-by-weight ratio that supported its candidacy as a
clean fuel. The obvious advantage of hydrogen is its environmentally favorable profile and the avail-
ability of multiple sources to produce hydrogen, such as water, methane, gasoline or coal, and so on.
Out of these sources, water seems to be an ultimate source of hydrogen because its decomposition
produced hydrogen and oxygen without releasing noxious carbon content in the air.
Several popular methods have been discussed for water splitting including the solar cleavage
of water using photocatalytic and photoelectrocatalytic techniques accompanied with their related
conceptual details and representative examples. But these techniques still need improvements to
meet required efficiency. Research into the development of an appropriate material continues for
large-scale hydrogen production with high efficiency. It is hoped that we will hear good news from
the photocatalytic world of materials. The efficiency of hydrogen production can be increased by
improving material engineering and systems design. In order to meet the ultimate target of achiev-
ing benchmark STH efficiency (i.e., more than 10% for the solar water splitting), one has to make
it commercially competitive and resolve the hurdles in the path of economical, green, and clean
production of hydrogen. Unlike past practices, where materials are discovered through trial and
error, we are going to possess more novel materials and rational device designs based on theory and
experimental simulations.
The hydrogen economy is proposed to suppress the negative effects of conventional fuels, which
release the load of the carbon into the atmosphere, and hydrogen is promoted as an environmen-
tally cleaner source of energy to end users (especially in transportation applications), to cut down
the release of air pollutants including carbon dioxide into the atmosphere at the point of end use.
Other technical obstacles include the issue of storage and the generation of highly pure (99.999%)
hydrogen that is required for fuel cells. Although hydrogen engine conversion technology is more
economical than fuel cells, the most significant drawback is its high production cost and difficulties
in storage during transport in comparison to traditional sources of energy. Future innovations in
technology and chemistry will help us to solve this issue. The moral of the story is that hydrogen
can be considered as a fuel of the future that still needs advancement in available technologies for
trouble-free storage and ample supply at low cost.
Hydrogen 169

REFERENCES
Abanades, S., and G. Flamant. 2006. Thermochemical hydrogen production from two step solar driven water
splitting cycle based on cerium oxides. Solar Energy. 80: 1611–1623.
Abdi, F. F., L. Han, A. H. M. Smets, M. Zeman, B. Dam, and R. Van de Krol. 2013. Efficient solar water
splitting by enhanced charge separation in a bismuth vanadate-silicon tandem photoelectrode. Nat.
Commun. 4: 2195. doi:10.1038/ncomms3195
Ahn, K. S., Y. Yan, S. Shet, T. Deutsch, J. Turner, and M. Al-Jassim. 2007. Enhanced photoelectrochemical
responses of ZnO films through Ga and N codoping. Appl. Phys. Lett. 91: 231909.
Ahn, K. S., Y. Yan, S. Shet, K. Jones, T. Deutsch, J. Turner, and M. Al-Jassim. 2008. ZnO nanocoral structures
for photoelectrochemical cells. Appl. Phys. Lett. 93: 163117.
Allam, N. K., and C. A. Grimes. 2009. Effect of rapid infrared annealing on the photoelectrochemical proper-
ties of anodically fabricated TiO2 nanotube arrays. J. Phys. Chem. C. 113: 7996–7999.
Arachchige, S. M., J. Brown, and K. J. Brewer. 2008. Photochemical hydrogen production from water using
the new photocatalyst [{(bpy)2Ru(dpp)2RhBr2](PF6)5 J. Photochem. Photobiol. A: Chem. 197: 13–17.
Arachchige, S. M., J. R. Brown, E. Chang, A. Jain, D. F. Zigler, et al. 2009. Design considerations for a system
for photocatalytic hydrogen production from water employing mixed-metal photochemical molecular
devices for photoinitiated electron collection. Inorg. Chem. 48: 1989–2000.
Ayers, W. 1983. US Patent 4,466,869 Photolytic production of hydrogen. November.
Balzani, V., A. Credi, and M. Venturi. 1997. Photoprocesses. Curr. Opin. Chem. Biol. 1: 506–513.
Bao, J. 2014. Recent developments in photocatalytic solar water splitting, Mater. Today. 17: 208–209.
Barnes, W. L., A. Dereux, and T. W. Ebbesen. 2003. Surface plasmon subwavelength optics. Nature. 424:
824–830.
Boettcher, S. W., J. M. Spurgeon, M. C. Putnam, E. L. Warren, D. B. Turner-Evans, M. D. Kelzenberg, et al.
2010. Energy-conversion properties of vapor-liquid-solid-grown silicon wire-array photocathodes.
Science. 327: 185–187.
Boettcher, S. W., E. L. Warren, M. C. Putnam, E. A. Santori, D. Turner- Evans, M. D. Kelzenberg, et al. 2011.
Photoelectrochemical hydrogen evolution using Si microwire arrays. J. Am. Chem. Soc. 133: 1216–1219.
Burns, L. D., J. B. McCormick, and E. E. Borroni-Bird. 2002. Vehicle of change. Sci. Am. 287: 64–73.
California Energy Commission. 2002. Fuel cells: Cost. California distributed energy resource guide.
January 18.
Charvin, P., S. Abanades, F. Lemort, and G. Flamant. 2008. Analysis of solar chemical processes for hydrogen
production from water splitting thermochemical cycles. Energy Convers. Manage. 49: 1547–1556.
Chen, H. M., C. K. Chen, C.-J. Chen, L.-C. Cheng, P. C. Wu, B. H. Cheng, et al. 2012. Plasmon induc-
ing effects for enhanced photoelectrochemical water splitting: X-ray absorption approach to electronic
structures. ACS Nano. 6: 7362–7372.
Chen, H., Y. Chen, H.-T. Hsieh, and N. Siegel. 2007. CFD Modeling of gas particle flow within a solid particle
solar receiver. ASME J. Solar Energy Eng. 129: 160–170.
Choi, W., A. Termin, and M. R. Hoffmann. 1994. The role of metal ion dopants in quantum-sized TiO2:
Correlation between photoreactivity and charge carrier recombination dynamics. J. Phys. Chem. 98:
13669–13679.
Chouhan, N., and R.-S. Liu. 2011. Electrochemical technologies for energy storage and conversion. In:
Electrochemical Technologies for Energy Storage and Conversion, Vol. 1, J. Zhang, L. Zhang, H. Liu,
A. Sun, and R.-S. Liu (Eds.), Wiley-VCH, Weinheim, p. 27.
Chouhan, N., C. L. Yeh, S. F. Hu, R. S. Liu, W. S. Chang, and K. H. Chen. 2011. Photocatalytic CdSe
QDs-decorated ZnO nanotubes: An effective photoelectrode for splitting water. Chem. Commun. 47:
3493–3495.
Crabtree, G. W., M. S. Dresselhaus, and M. V. Buchanan. 2004. The hydrogen economy. Phys. Today. 57: 39.
Das, C., P. Roy, M. Yang, H. Jha, and P. Schmuki. 2011. Nb-doped TiO2 nanotube for enhanced photoelectro-
chemical water-splitting. Nanoscale. 3: 3094–3096.
Dincer, I., and A. S. Joshi. 2013. Solar Based Hydrogen Production System. Springer, New York, pp. 27–69.
Dincer, I., and C. Zamfirescu. 2012. Sustainable Energy Systems and Application. New York: Springer,
pp. 519–632.
Du, P., J. Schneider, F. Li, W. Zhao, U. Patel, F. N. Castellano, and R. Eisenberg. 2008. Bi- and terpyridyl
platinum(II) chloro complexes: Molecular catalysts for the photogeneration of hydrogen from water or
simply precursors for colloidal platinum? J. Am. Chem. Soc. 130: 5056–5058.
Elvington, M., and K. J. Brewer. 2006. Photoinitiated electron collection at a metal in a rhodium-centered
mixed-metal supramolecular complex. Inorg. Chem. 45: 5242–5244.
170 Solar Energy Conversion and Storage

Elvington, M., J. Brown, S. M. Arachchige, and K. J. Brewer. 2007. Photocatalytic hydrogen production from
water employing a Ru, Rh, Ru molecular device for photoinitiated electron collection. J. Am. Chem. Soc.
129: 10644–10645.
Fei, H., R. Ye, G. Ye, Y. Gong, Z. Peng, Z. Fan, et al. 2014. Boron- and nitrogen-doped graphene quan-
tum dots/graphene hybrid nanoplatelets as efficient electrocatalysts for oxygen reduction. ACS Nano.
8: 10837–10843.
Fujishima, A., and K. Honda. 1972. Electrochemical photolysis of water at a semiconductor electrode. Nature.
238: 37–38.
Fujishima, A., and D. A. Tryk. 2012. Energy carriers and conversion systems: Vol. 1. Photochemical and pho-
toelectrochemical water splitting. In: Encyclopedia of Life Support Systems, T. Ohta and T. N. Vezirozeu
(Eds.). ELOSS & UNESCO.
Grimes, C. A., and G. K. Mor. 2009. TiO2 Nanotube Arrays: Synthesis, Properties, and Applications. Springer,
New York.
Gurunathan, K., P. Maruthamuthu, and M. V. C. Sastri. 1997. Photocatalytic hydrogen production by dye-sensitized
Pt/SnO2 and Pt/SnO2/RuO2 in aqueous methyl viologen solution. Int. J. Hydrogen Energy. 22: 57–62.
Hairston, D. 1996. Hail hydrogen. Chem. Eng. 103: 59–62.
Hamedani, H. A., N. K. Allam, H. Garmestani, and M. A. El-Sayed. 2011. Electrochemical fabrication of
strontium-doped TiO2 nanotube array electrodes and investigation of their photoelectrochemical prop-
erties. J. Phys. Chem. C. 115: 13480–13486.
Ingram, D. B., P. Christopher, J. L. Bauer, and S. Linic. 2011. Predictive model for the design of plasmonic
metal/semiconductor composite photocatalysts. ACS Catal. 1: 1441–1447.
Jing, D., and L. Guo. 2007. WS2 sensitized mesoporous TiO2 for efficient photocatalytic hydrogen production
from water under visible light irradiation. Catal. Commun. 8: 795–799.
Kato, H., K. Asakura, and A. Kudo. 2003. Highly efficient water splitting into H and O over lanthanum-doped
NaTaO photocatalysts with high crystallinity and surface nanostructure. J. Am. Chem. Soc. 125: 3082.
Kibsgaard, J., T. F. Jaramillo, and F. Besenbacher. 2014. Building an appropriate active-site motif into a
hydrogen-evolution catalyst with thiomolybdate [Mo3S13]2 clusters. Nat. Chem. 6: 248–253.
Kolb, G. J., N. P. Siegel, and R. B. Diver. 2007. Central-station solar hydrogen power plant, ASME J. Solar
Energy Eng. 129: 179–183.
Kudo, A., and Y. Miseki. 2009. Heterogeneous photocatalyst materials for water splitting. Chem. Soc. Rev.
38: 253–278.
Kurihara, T., H. Okutomi, Y. Miseki, H. Kato, and A. Kudo. 2006. Highly efficient water splitting over
K3Ta3B2O12 photocatalyst without loading cocatalyst. Chem. Lett. 35: 274–275.
Lal, S., S. Link, and N. Halas. 2007. Nano-optics from sensing to waveguiding, Nat. Photonics. 1: 641–648.
Lei, P., M. Hedlund, R. Lomoth, H. Rensmo, O. Johansson, and L. Hammarström. 2008. The role of colloid
formation in the photoinduced H2 production with a RuII−PdII supramolecular complex: A study by
GC, XPS, and TEM. J. Am. Chem. Soc. 130: 26–27.
Liao, L., Q. Zhang, Z. Su, Z. Zhao, Y. Wang, Y. Li, et al. 2014. Efficient solar water-splitting using a nanocrys-
talline CoO photocatalyst. Nat. Nanotechnol. 9: 69–73.
Logan, B. E., and J. M. Regan. 2006. Electricity-producing bacterial communities in microbial fuel cells.
Trend. Microbiol. 14: 512–518.
Luan, P., M. Xie, X. Fu, Y. Qu, X. Sun, and L. Jing. 2015. Improved photoactivity of TiO2–Fe2O3 nanocom-
posites for visible-light water splitting after phosphate bridging and its mechanism. Phys. Chem. Chem.
Phys. 17: 5043–5050.
Luo, J., J.-H. Im, M. T. Mayer, M. Schreier, M. K. Nazeeruddin, N.-G. Park, et al. 2014. Water photolysis at
12.3% efficiency via perovskite photovoltaics and Earth-abundant catalysts. Science. 345: 1593–1596.
Ma, S. S. K., T. Hisatomi, K. Maeda, Y. Moriya, and K. Domen. 2012. Enhanced water oxidation on Ta3N5
photocatalysts by modification with alkaline metal Salts. J. Am. Chem. Soc. 134: 19993–19996.
Ma, S. S. K., K. Maeda, T. Hisatomi, M. Tabata, A. Kudo, and K. Domen. 2013. A redox-mediator-eree solar-
driven Z-scheme water-splitting system consisting of modified Ta3N5 as an oxygen-evolution photocata-
lyst. Chem. Eur. J. 19: 7480–7486.
Ma, G., T. Minegishi, D. Yokoyama, J. Kubota, and K. Domen. 2011. Photoelectrochemical hydrogen production
on Cu2ZnSnS4/Mo-mesh thin-film electrodes prepared by electroplating. Chem. Phys. Lett. 501: 619–622.
Maeda, K., and K. Domen. 2010. Photocatalytic water splitting: Recent progress and future challenges.
J. Phys. Chem. Lett. 1: 2655–2661.
Maeda, K., M. Higashi, D. Lu, R. Abe, and K. Domen. 2010. Efficient nonsacrificial water splitting through
two-step photoexcitation by visible light using a modified oxynitride as a hydrogen evolution photocata-
lyst. J. Am. Chem. Soc. 132: 5858–5868.
Hydrogen 171

Maeda, K., K. Teramura, and K. Domen. 2008. Effect of post-calcination on photocatalytic activity of
(Ga1–xZn x)(N1–xOx) solid solution for overall water splitting under visible light. J. Catal. 254: 198–204.
Maier, S. A. 2007. Plasmonics Fundamentals and Applications. Springer, New York, pp. 21–34, 65–88.
Maier, S. A., and H. A. Atwater. 2005. Plasmonics: Localization and guiding of electromagnetic energy in
metal/dielectric structures. J. Appl. Phys. 98: 011101.
McAlister, R. 1999. What causes the Hindenberg disaster? Hydrogen Today. 10: 6–8.
Minegishi, T., N. Nishimura, J. Kubota, and K. Domen. 2013. Photoelectrochemical properties of LaTiO2N
electrodes prepared by particle transfer for sunlight-driven water splitting. Chem. Sci. 4: 1120–1124.
Moriya, M., T. Minegishi, H. Kumagai, M. Katayama, J. Kubota, and K. Domen. 2013. Stable hydrogen evolu-
tion from CdS-modified CuGaSe2 photoelectrode under visible-light irradiation. J. Am. Chem. Soc. 135:
3733–3735.
Mubeen, S., J. Lee, N. Singh, S. Krämer, G. D. Stucky, and M. Moskovits. 2013. An autonomous photosyn-
thetic device in which all charge carriers derive from surface plasmons. Nat. Nanotechnol. 8: 247–251.
Murphy, A. B. 2008. Does carbon doping of TiO2 allow water splitting in visible light? Comments on nano-
tube enhanced photoresponse of carbon modified (CM)-n-TiO2 for efficient water splitting. Sol. Energy
Mater. Solar Cells. 92: 363–367.
Murphy, A. B., P. R. F. Barnes, L. K. Randeniya, I. C. Plumb, I. E. Grey, M. D. Horne, et al. 2006. Efficiency
of solar water splitting using semiconductor electrodes. Int. J. Hydrogen. Energy. 31: 1999–2017.
Naess, S. N., A. Elgsaeter, G. Helgesen, and K. D. Knudsen. 2009. Carbon nanocones: Wall structure and
morphology. Sci. Technol. Adv. Mater. 10: 065002.
Ni, M., M. K. H. Leung, D. Y. C. Leung, and K. Sumathy. 2007. A review and recent developments in pho-
tocatalytic water-splitting using TiO2 for hydrogen production. Renew. Sust. Energy Rev. 11: 401–425.
NREL. 2009. Current state-of-the-art hydrogen production cost estimate using water electrolysis. NREL/
BK-6A1-46676 (http://1.usa.gov/VFOj30), pp. 23–24.
Ogden, J. M. 1999. Prospects for building a hydrogen energy infrastructure. Ann. Res. Energy Environ. 24: 227–279.
Ozawa, H., M. Haga, and K. Sakai. 2006. A photo-hydrogen-evolving molecular device driving visible-light-
induced EDTA-reduction of water into molecular hydrogen. J. Am. Chem. Soc. 128: 4926–4927.
Ozawa, H., Y. Yokoyama, M. Haga, and K. Sakai. 2007. Syntheses, characterization, and photo-hydrogen-
evolving properties of tris(2,2′-bipyridine)ruthenium(II) derivatives tethered to a cis-Pt(II)Cl2 unit:
Insights into the structure-activity relationship. Dalton Trans. 1197–1206.
Pauling, L. 1970. General Chemistry, Section 15-2. San Francisco, CA, Dover.
Perkins, C., and A. W. Weimer. 2004. Likely near-term solar-thermal water splitting technologies. Int. J.
Hydrogen Energy. 29: 1587–1599
Perret, R., W. Alan. G. Besenbruch, R. Diver, M. Lewis, Y. Chen, and K. Roth. 2007. Development of solar-
powered thermochemical production of hydrogen from water, DOE Hydrogen Program FY Annual
Progress Report, pp. 128–135.
Pirhashemi, M., and A. H.-Yangjeh. 2014. Preparation of AgCl–ZnO nanocomposites as highly efficient visi-
ble-light photocatalysts in water by one-pot refluxing method. J. Alloys Compds. 601: 1–8.
Rau, S., B. Schäfer, D. Gleich, E. Anders, M. Rudolph, M. Friedrich, et al. 2006. A supramolecular photocatalyst for
the production of hydrogen and the selective hydrogenation of tolane. Angew. Chem. Int. Ed. 45: 6215–6218.
Ritchie, R. H. 1957. Plasma losses by fast electrons in thin films. Phys. Rev. 106: 874–881.
Rosenbaum, M., Z. He, and L. T. Angenent. 2010. Light energy to bioelectricity: Photosynthetic microbial fuel
cells. Curr. Opin. Biotechnol. 21: 259–264.
Sauve, G., M. E. Cass, G. Coia, S. J. Doig, I. Lauermann, K. E. Pomykal, N. S. Lewis, et al. 2000. Dye sensiti-
zation of nanocrystalline titanium dioxide with osmium and ruthenium polypyridyl complexes. J. Phys.
Chem. B. 104: 6821–6836.
Scholz, W. 1993. Processes for industrial production of hydrogen and associated environmental effect. Gas
Sep. Purif. 7: 131.
Sharmoukh, W., and N. K. Allam. 2012. TiO2 Nanotube-based dye-sensitized solar cell using new photosen-
sitizer with enhanced open-circuit voltage and fill factor. ACS Appl. Mater. Interfaces. 4: 4413–4418.
Singh, G. P., K. M. Shrestha, A. Nepal, K. J. Klabunde, and C. M. Sorensen. 2014. Graphene supported plas-
monic photocatalyst for hydrogen evolution in photocatalytic water splitting, Nanotechnology. 25: 265701.
Stempel, T., M. Aggour, K. Skorupska, A. Munoz, and H. J. Lewerenz. 2008. Efficient photoelectrochemical
nanoemitter solar cell. Electrochem. Commun. 10: 1184–1186.
Sykora, M., K. Maruszewski, S. M. Treffert-Ziemelis, and J. R. Kincaid. 1998. A synthetic strategy for the
construction of zeolite-entrapped organized molecular assemblies. Preparation and photophysical char-
acterization of interacting adjacent cage dyads comprised of two polypyridine complexes of Ru(II).
J. Am. Chem. Soc. 120: 3490–3498.
172 Solar Energy Conversion and Storage

Szklo, A., and R. Schaeffer. 2007. Fuel specification, energy consumption and CO2 emission in oil refineries.
Energy. 32: 1075–1092.
Tachibana, Y., L. Vayssieres, and J. R. Durrant. 2012. Artificial photosynthesis for solar water-splitting. J. Nat.
Photonics. 6: 511–518.
Tao, Y., Y. Chen, Y. Wu, Y. He, and Z. Zhou. 2007. High hydrogen yield from a two-step process of dark-and
photo-fermentation of sucrose, Int. J. Hydrogen Energy. 32: 200–206.
Tester, J. W., E. M. Drake, M. J. Driscoll, M. W. Golay, and W. A. Peters. 2005. Sustainable Energy: Choosing
among Options. Cambridge, MA: MIT Press.
Torimoto, T., H. Horibe, T. Kameyama, K. Okazaki, S. Ikeda, M. Matsumura, et al. 2011. Plasmon-enhanced
photocatalytic activity of cadmium sulfide nanoparticle immobilized on silica-coated gold particles.
J. Phys. Chem. Lett. 2: 2057–2062.
Yamada, S., A. Y. Nosaka, and Y. Nosaka. 2005. Fabrication of US photoelectrodes coated with titania
nanosheets for water splitting with visible light. J. Electroanal. Chem. 585: 105–112.
Yang, X. Y., A. Wolcott, G. M. Wang, A. Sobo, R. C. Fitzmorris, F. Qian, et al. 2009. Nitrogen-doped ZnO
nanowire arrays for photoelectrochemical water splitting. Nano Lett. 9: 2331–2336.
Yeh, T. F., C. Y. Teng, S. J. Chen, and H. Teng. 2014. Nitrogen-doped graphene oxide quantum dots as photo-
catalysts for overall water-splitting under visible light illumination. Adv. Mater. 26: 3297–3303.
Yokoyama, D., T. Minegishi, K. Jimbo, T. Hisatomi, G. Ma, M. Katayama, et al. 2010. H2 Evolution from
water on modified Cu2ZnSnS4 photoelectrode under solar light. Appl. Phys. Express. 3: 101202.
Zeng, K., and D. Zhang. 2010. Recent progress in alkaline water electrolysis for hydrogen production and
applications. Prog. Energy Combust. Sci. 36: 307–326.
Zhang, C., Y. Z. Jia, Y. Jing, Y. Yao, J. Maa, and J. Sun. 2013. Effect of non-metal elements (B, C, N, F, P, S)
mono-doping as anions on electronic structure of SrTiO3. Comp. Mater. Sci. 79: 69–74.
Zhang, X., Y. Li, J. Zhao, S. Wang, Y. Li, H. Dai, and X. Sun. 2014. Advanced three-component ZnO/Ag/CdS
nanocomposite photoanode for photocatalytic water splitting. J. Power Sources. 269: 466–472.
Zittel, W., and R. Wurster. 1996. Production of hydrogen. In Part 4: Production from electricity by means
of electrolysis. HyWeb: Knowledge—Hydrogen in the Energy Sector. Ludwig-Bölkow-Systemtechnik
GmbH.
9 Photocatalytic Reduction
of Carbon Dioxide
Guoqing Guan, Xiaogang Hao, and Abuliti Abudula

CONTENTS
9.1 Introduction........................................................................................................................... 173
9.2 Characteristics of CO2 Molecule........................................................................................... 174
9.3 Mechanism of Photocatalytic Reduction of CO2................................................................... 174
9.4 Photocatalysts for Reduction of CO2..................................................................................... 177
9.4.1 TiO2 and Related Titanium-Containing Solids......................................................... 177
9.4.2 Other Metal Oxide Photocatalysts............................................................................. 179
9.4.3 Nonoxide Semiconductor Photocatalysts.................................................................. 179
9.5 Operation Conditions............................................................................................................. 181
9.5.1 Pressure Effect........................................................................................................... 181
9.5.2 Reductant Effect........................................................................................................ 181
9.5.3 Temperature Effect.................................................................................................... 181
9.5.4 Reactor-Type Effect................................................................................................... 182
9.6 Conclusions and Outlook....................................................................................................... 182
References....................................................................................................................................... 183

9.1 INTRODUCTION
Carbon dioxide (CO2) is the chief greenhouse gas that results from human activities, which causes
global warming and climate change. Based on the atmospheric CO2 measurement data at Mauna
Loa Observatory, Hawaii, the concentration of CO2 in the atmosphere is increasing each decade at
an accelerating rate (ESRL 2014). The upper safety limit for atmospheric CO2 is 350 ppm; however,
atmospheric CO2 levels have remained higher than 350 ppm since early 1988, now reaching levels
of 399 ppm. There are three possible strategies for reducing atmospheric CO2:

• Reduction of the produced amount of CO2 from power generation stations and industrial
sites
• Usage of CO2 as a feedstock of nearly zero cost for conversion to fuels and chemicals
• Storage of it using carbon sequestration and storage (CCS) technologies (Centi and
Parathoner 2009; Mikkelsen et al. 2010; Richter et al. 2013; Hu et al. 2013)

The conversion of CO2 to fuels, chemicals, and materials has attracted much attention in recent years
(Ma et al. 2009; Morris et al. 2009; Jensen et al. 2011; Windle and Perutz 2012; Liu et al. 2012; Inglis
et al. 2012; Tran et al. 2012; Dhakshinamoorthy et al. 2012; Mori et al. 2012; Fan et al. 2013; Izumi
2013; Handoko et al. 2013; Tahir and Amin 2013; Stechel and Miller 2013; Mao et al. 2013;
Kondratenko et al. 2013; Das and Daud 2014). However, CO2 is a kinetically and thermodynami-
cally stable compound, which is difficult to oxidize or reduce to other fuels and/or chemicals at low
temperatures. Reduction of CO2 needs high energy input, cofeeding of a high-energy reactant such
as H2, and an excellent catalyst capable of driving its selective conversion to targeted chemicals.
CO2 can be theoretically reduced to hydrocarbons and/or alcohols at high temperature and high

173
174 Solar Energy Conversion and Storage: Photochemical Modes

pressure. However, the process is quite complex, and a large amount of energy is essential. It would
be attractive to capture CO2 from the atmosphere or the exhaust of power stations or factories, and
convert it to fuel and chemicals by using photocatalysts and a sustainable source of energy such
as sunlight. In combination with photocatalytic H2 production from water, the solar-driven CO2
reduction to fuels and chemicals could be a very promising process for utilization of CO2. Here, the
mechanism of photocatalytic reduction of CO2, the catalytic activity of the developed catalysts for
the CO2 reduction, and its advantages and challenges are introduced and discussed.

9.2  CHARACTERISTICS OF CO2 MOLECULE


The CO2 molecular geometry is linear in its ground state with two polar C=O bonds, in which oxy-
gen atoms show a weak Lewis basicity, while the carbon atom is electrophilic (Centi and Parathoner
2009; Mikkelsen et al. 2010). Because it is centrosymmetric, the molecule has no electrical dipole,
and only two vibrational bands are observed in the infrared (IR) spectrum: an antisymmetric
stretching mode at 2349 cm−1 and a bending mode near 666 cm−1. There is also a symmetric stretch-
ing mode at 1388 cm−1 in its Raman spectrum. The CO2 is soluble in water, where it reversibly forms
weak acid H2CO3, and easily reacts with basic compounds. The reactions of CO2 are dominated
by nucleophilic attack at the carbon atom, and it reacts with electron-donating reagents. Carbon
dioxide has low reactivity, but it can be activated from nucleophilic centers on the surfaces of solids
or in organometallic complexes. However, it is necessary to overcome a thermodynamic barrier by
providing the energy for the reaction. In other words, thermochemical reduction of CO2 is always
an endothermic reaction, and it has to be realized at high temperatures even in the presence of cata-
lysts. For instance, the reduction of CO2 by H2O to form methanol and methane requires 726 and
802 kJ mol−1 thermal energy, respectively.

9.3  MECHANISM OF PHOTOCATALYTIC REDUCTION OF CO2


Photocatalytic reduction of CO2 with a reducing agent such as H2 and H2O into useful fuels and
chemicals such as CO, HCOOH, HCHO, CH4, CH3OH, and C2H5OH has been studied for decades
(Centi and Parathoner 2009; Ma et al. 2009; Morris et al. 2009; Mikkelsen et al. 2010; Jensen et al.
2011; Windle and Perutz 2012; Liu et al. 2012; Inglis et al. 2012; Tran et al. 2012; Dhakshinamoorthy
et al. 2012; Mori et al. 2012; Richter et al. 2013; Hu et al. 2013; Fan et al. 2013; Izumi 2013; Handoko
et al. 2013; Tahir and Amin 2013; Stechel and Miller 2013; Mao et al. 2013; Kondratenko et al. 2013;
Das and Daud 2014). Direct photolysis of CO2 to CO and O2 in the absence of a reducing agent is
also possible (Corma and Garcia 2013). Carbon dioxide can only absorb the photons below 200 nm,
which belongs to the deep UV spectral region:

2CO 2 + hν (<200 nm ) → 2CO + O2    ΔH = + 257 kJ mol−1 (9.1)


In order to get the light in this region, pure quartz or metal halides and vacuum are required.
However, the key issues of the direct photolysis of CO2 are the efficiency and energy consumption
per mole of CO2 conversion. It is estimated that the energy consumption value is about 0.3 GJ mol−1
CO2 for this process. This energy consumption is high if the efficiency of the electricity to deep UV
light conversion for mercury lamps is considered. As a sustainable source of energy, only a small
amount of deep UV light (100–280 nm) is present in the UV light that reaches Earth. How to use a
large amount of other photons included in the solar insolation for the CO2 reduction should be one
of the main challenges in the field of photocatalytic reduction of CO2.
The reduction of CO2 by one electron to form a carbon dioxide anion radical (CO2−•) is highly
unfavorable, requiring a highly negative potential of –1.9 V versus NHE (Figure 9.1) (Morris et al.
Photocatalytic Reduction of Carbon Dioxide 175

–2 CO2/CO2
SiC
Cu2O CB
ZnS
GaP HCOOH/CO2
CdS
–1 CO/CO2

2.0 eV
TiO2 ZnO HCHO/CO2
WO3 H2/H2O

Potential vs. NHE (V)


0 Fe2O3 CH3OH/CO2

3.0 eV
3.6 eV
CH4/CO2

2.4 eV

2.3 eV
O2/H2O
1
2.8 eV

3.2 eV

2.1 eV
3.2 eV

Redox potentials
VB (pH = 7)

FIGURE 9.1  Band-edge positions of typical semiconductors and redox potentials at pH = 7 for CO2 reduction.

2009; Corma and Garcia 2013). Rapid reduction requires an overpotential of at least 0.6 V due to
the kinetic restrictions imposed by the structural difference between linear CO2 and bent CO2−•.
In the photocatalysis process, electrons are provided by a semiconductor exposed to light (Inoue
et al. 1979). The photocatalysis over a semiconductor is initiated by the adsorption of photons in the
light with equal or higher energy than the bandgap energy of the semiconductor. The excitation of
the electron from the valence band (VB) to the conduction band (CB) results in electron vacancy
called a hole in the VB, and the electron–hole pair is generated in the semiconductor (Equation 9.2)
(Figure 9.2).
Thereafter, the separation and migration of these photogenerated electrons and holes to the surface
sites for the oxidation of H2O (donor) by holes to O2 (Equation 9.3) and the reduction of H2O or CO2
(acceptor) by electrons to H2 or CO and other organic compounds become important (Equations 9.4
through 9.10). Otherwise, the photogenerated electron–hole pairs will recombine. The recombina-
tion time of the electron–hole pairs is of the order of 10 −9 s, whereas the chemical interaction with
the adsorbed species has a longer time of 10 −8 to 10 −3 s. Therefore, it is fundamentally important to

H2O CO + HO2
(Acceptor) H2 (Acceptor)

CO, HCOOH
HCHO, CH3OH
e CH4...
CB

Recombination
VB

H2O h+
(Donor)
O2

FIGURE 9.2  The photocatalytic reduction of CO2 on the catalyst surface.


176 Solar Energy Conversion and Storage: Photochemical Modes

avoid the recombination by improving their separation and migration to the surface sites during the
photocatalysis process:

Photocatalyst + hν (≥ Band gap) → e− + h+ (9.2)

2H 2O + 4h → O 2 + 4H+ → E 0 versus NHE = +0.81 V (9.3)

2H + + 2e – → H 2 → E 0 versus NHE = – 0.42 V (9.4)

CO 2 + e− → CO 2 −• → E 0 versus NHE = – 1.90 V (9.5)

CO 2 + 2H+ + 2e− → CO + H 2O → E 0 versus NHE = –0.53 V (9.6)

CO 2 + 2H+ + 2e− → HCOOH → E 0 versus NHE = –0.61 V (9.7)

CO 2 + 4H+ + 4e− → HCHO + H 2O → E 0 versus NHE = –0.48 V (9.8)

CO 2 + 6H+ + 6e− → CH 3OH + H 2O → E 0 versus NHE = –0.38 V (9.9)

CO 2 + 8H+ + 8e− → CH 4 + 2H 2O → E 0 versus NHE = –0.24 V (9.10)

Thermodynamically, proton-coupled electrochemical reactions are possible at much less negative


potentials. Thus, photocatalytic reduction of CO2 to either CH3OH or CH4 seems to be more fea-
sible than reduction to CO2−•. However, the lower oxidation states of carbon in both are achieved
through kinetically unfavorable routes, because they require the transfer of six and eight electrons
to each carbon center, respectively, along with complex chemical steps involving proton transfers
(Inglis et al. 2012). It should be noted that the standard reduction potential of H2O to produce H2 is
much lower than that of CO2 to generate CO2−•. Thus, in the case of the photocatalytic reduction of
CO2 with H2O, the photocatalytic reduction of H2O to generate molecular hydrogen at a much lower
reduction potential will compete with the reduction of CO2. Therefore, this thermodynamic and
kinetic contradiction should be considered by creating spatially separated centers for electrons and
protons to avoid their collapse before being transferred to CO2.
The photocatalytic reduction of CO2 using water as the reducing agent always results in a small
amount of organic products (Tahir and Amin 2013). Several factors can be considered:

• The solubility of CO2 at pH = 7 is very low so that it is more difficult to contact the catalyst
surface than H2O. Photocatalytic reduction of CO2 can be performed in alkaline solution to
solve this problem. The formation of CO32− and HCO3− in the alkaline solution can acceler-
ate the CO2 reduction. Furthermore, the OH− in the alkaline solution can serve as a strong
hole scavenger to promote the photocatalytic activity (Mao et al. 2013).
• H2O is a poor electron donor. In this case, it is possible to improve the efficiency by employ-
ing other sacrificial electron donors such as isopropyl alcohol, dimethylformamide, and
phosphoric acid (Mikkelsen et al. 2010, Dhakshinamoorthy et al. 2012).
• Photocatalytic activity of the developed semiconductors is low; hence, in addition to photo-
reduction mechanisms, other factors should be considered for a practical process.
Photocatalytic Reduction of Carbon Dioxide 177

9.4  PHOTOCATALYSTS FOR REDUCTION OF CO2


Many traditional semiconductors such as TiO2, CdS, Fe2O3, WO3, ZnO, ZrO2, and their various
combinations have been tested for the photoreduction of CO2 (Centi and Parathoner 2009; Ma et al.
2009; Morris et al. 2009; Mikkelsen et al. 2010; Jensen et al. 2011; Windle and Perutz 2012; Liu
et al. 2012; Inglis et al. 2012; Tran et al. 2012; Dhakshinamoorthy et al. 2012; Mori et al. 2012;
de Richter et al. 2013; Hu et al. 2013; Fan et al. 2013; Izumi 2013; Handoko et al. 2013; Tahir and
Amin 2013; Stechel and Miller 2013; Mao et al. 2013; Kondratenko et al. 2013; Das and Daud 2014).
Recently, some new semiconductors such as copper chalcopyrite compounds, CuInSe2, CuInS2, and
CuInGaSe2, and carbon nitride (g-C3N4) were also considered as the photocatalysts for the photo-
reduction of CO2 (Tran et al. 2012; Kondratenko et al. 2013). In general, as a good photocatalyst,
these should have high catalytic activity with high stability and durability, with a low tendency to
photocorrosion.

9.4.1  TiO2 and Related Titanium-Containing Solids


Inoue et al. (1979) first reported photocatalytic reduction of CO2 to organic compounds including
HCOOH, CH3OH, and/or HCHO by irradiation of aqueous saturated CO2 using a Xe lamp in the
presence of various semiconductors such as TiO2, ZnO, and WO3. Since then, the study on artificial
photosynthesis boomed with typical focus on TiO2. Comparing to other semiconductors, TiO2 is
relatively low cost, nontoxic, and fairly stable. Three phases of TiO2 (i.e., rutile, anatase, and brook-
ite) exist naturally. Anatase and rutile phases reveal definite bandgap energy at about 3.2 and 3.0 eV,
respectively. The best photoactivity can be achieved by combining anatase with a slight amount of
rutile (Das and Daud 2014). Pure TiO2 in its anatase form shows good catalytic performance when
it is irradiated with photons of wavelength shorter than 380  nm. Generally, P25, a kind of TiO2
made by Degussa with a particle size of about 30 nm, containing rutile, is considered as the refer-
ence material for the photocatalytic applications of TiO2 and other developing photocatalysts. The
photocatalytic activity of TiO2 can be enhanced by

• Control of its particle morphology by decreasing its size to nanolevel, synthesizing titania
nanotube or nanorod, and supporting aluminosilicates
• Doping with metallic or nonmetallic elements
• Photosensitization

Photoreduction of CO2 by using TiO2 nanoparticles with size in the range of 4.5–29 nm were inves-
tigated by Kočí et al. (2009). As the nanoparticle size decreased, higher yields of CH3OH and CH4
were obtained. The optimum particle size corresponding to the highest yields of both products
(CH3OH: 1.15  μmol/g-catalyst; CH4: 9.5  μmol/g-catalyst) was found to be 14  nm. Theoretically,
the bandgap between the CB and VB becomes larger as the TiO2 particle size decreases, making it
suitable and applicable for the reduction of CO2 (Mori et al. 2012). However, further decreasing the
particle size resulted in a decrease in the catalytic activity due to the competing effects of specific
surface area, charge–carrier dynamics, and light absorption efficiency.
The TiO2 in the form of a nanotube or a nanorod also exhibits higher photocatalytic activity as
compared to P25. The electron transport along the long axes of the crystallites of nanotube and
nanorod is very fast, and it is beneficial for the charge separation. Furthermore, the light can be
trapped inside the nanotube or nanorod, not only by the light scattering but also by the crystal pho-
tonic effect that appears when the wavelength of the photon matches the distances between the TiO2
nanotubes or nanorods (Dhakshinamoorthy et al. 2012).
Titanium oxide moieties implanted and isolated in the silica matrices of zeolite and mesoporous
silica materials at the atomic level, which have been called single-site photocatalysts, have also been
tested for the photoreduction of CO2 (Mori et al. 2012). In this case, the isolated Ti atoms can be
178 Solar Energy Conversion and Storage: Photochemical Modes

O2– O–

Ti4+ Ti3+

O2– O2– O2– O2– O2– O2–

Isolated Ti-oxide Excited state

FIGURE 9.3  The electronic state change in isolated Ti-oxide molecular species.

substituted with Si atoms in the silica matrices (Si/Ti > 30) and coordinated tetrahedrally with oxygen
atoms. If it is excited by UV irradiation, an electron motivation from the oxygen (O2−) to Ti4+ ions will
occur, and pairs of trapped hole centers (O−) and electron centers (Ti3+) are formed (Figure 9.3). Such
excited electron–hole pairs locate quite near each other compared to the electron–hole pairs in the bulk
TiO2, and unique photocatalytic activity could be obtained for the photoreduction of CO2. The results
obtained by using micro- and mesoporous titanosilicates such as TS-1, Ti-MCM-41, Ti-MCM-48,
Pt-Ti-MCM-48, titanium beta-zeolites (Ti-β) of hydrophilic Ti-β-OH and hydrophobic Ti-β-F, and
Ti-FSM-16 indicated that Ti-MCM-48 exhibits the highest reactivity and selectivity for the formation
of CH3OH due to the combined contribution of a high dispersion of titanium oxide moieties, large pore
size, and tridirectional channel structure. Pt doping in Ti-MCM-48 results in greater yield of CH4 and
a decrease in the CH3OH yield (Dhakshinamoorthy et al. 2012, Mori et al. 2012).
Ti-β-OH shows five times higher reactivity compared to Ti-β-F for the formation of CH4, but
Ti-β-F shows higher selectivity for the formation of CH3OH due to the different abilities of zeolite
pores on the H2O affinity.
TiO2 with a wide bandgap mainly absorbs ultraviolet photons with wavelengths shorter than
380 nm. However, solar light contains about 5% only of ultraviolet photons, and a large amount of
light in nature is visible light. Many efforts have been made to extend the spectral response of pure
TiO2 to visible light (Fujishima et al. 2008; Dhakshinamoorthy et al. 2012; Mori et al. 2012):

• Doping metal to replace Ti atoms in the lattice for introducing empty orbitals at energies
below the CB of pure TiO2 (Figure 9.4a)
• Coupling TiO2 with narrow-bandgap semiconductors
• Preparing oxygen-deficient TiO2
• Doping nonmetal atoms such as N, C, S, B, P, and F in the lattice to replace O atoms for
introducing occupied orbitals above the energy of the VB of pure TiO2 (Figure 9.4b)

(a) (b)

CB CB

VB VB

Metal doping Nonmetal doping

FIGURE 9.4  Change in bandgap structure by (a) replacing the Ti atom with another metal element in the
TiO2 lattice or (b) by replacing some O atoms in the TiO2 lattice by other nonmetallic species.
Photocatalytic Reduction of Carbon Dioxide 179

However, it is found that metal doping always results in photocorrosion phenomenon during long-
term operation and affects the stability and durability of the photocatalyst. In contrast, nonmetal
doping renders more durable and stable photocatalysts working in visible light. Among nonmetal
doping, N-doped TiO2 is the most effective in narrowing the bandgap.
Another method of metal doping is doping the metal on the surface of TiO2. Cu doping was
found to be the most efficient way to enhance the selectivity toward CH3OH production and for-
mation of formic acid derivatives (Tseng et al. 2002, 2004; Wu et al. 2005). It was found that Cu+
species on TiO2 plays a significant role in the photoelectrochemical production of CH3OH from
CO2 and H2. Cu metal can serve as an electron trapper, and it prohibits the recombination of elec-
tron and hole, thus, significantly increasing the photoefficiency. In contrast, Pt or Pd doping favors
selectivity toward CH4 and CO (Dhakshinamoorthy et al. 2012). It should be noted that the doping
amount has a great effect on catalytic activity. A third effect in the case of metal nanoparticle dop-
ing, especially for Au doping, is that a strong surface plasmon band in the visible light region can
be observed. In this case, electrons expelled from the Au nanoparticle will have enough energy to
move to CB in TiO2, even when irradiated by visible light (Dhakshinamoorthy et al. 2012; Mori
et al. 2012).

9.4.2  Other Metal Oxide Photocatalysts


In addition to TiO2, other metal oxide semiconductors can be used as photocatalysts for the reduc-
tion of CO2. Different metal oxides possess different bandgap energies and different CB and VB
positions versus NHE (Figure 9.1). It was found that the yield of methanol increases as the CB is
located more negative with respect to the redox potential of CH3OH/CO2. For instance, methanol
cannot be produced over WO3 catalyst, as its conduction band is more positive than the redox
potential of CH3OH/CO2. However, methane can be generated from it, while it has low bandgap
energy (2.8 eV) with respect to visible light irradiation (Figure 9.1). In order to improve the cata-
lytic activity, various synthesis methods such as sol-gel, hydrothermal, and solid-state reaction
methods are developed to get high-performance metal oxide semiconductors with some special
structures for the photoreduction of CO2. For instance, W18O49 nanowires, each with a diameter
of about 0.9 nm and with a large number of oxygen vacancies, were synthesized, and these show
potential for the photoreduction of CO2 in the visible light region (Xi et al. 2012). Hydrothermally
prepared monoclinic BiVO4 catalysts show higher ethanol yields (110  μmol/h/g-catalyst) than
tetragonal BiVO4, as the anchored CO32− to the Bi3+ sites on the external surface can receive the
photogenerated electrons effectively via a weak Bi-O bond from V 3d-block bands of BiVO3 (Liu
et al. 2012). Table 9.1 summarizes some typical metal oxide photocatalysts for the reduction of CO2
with relatively high yields.

9.4.3 Nonoxide Semiconductor Photocatalysts


Nonoxide semiconductor photocatalysts always have low bandgap energy so that they can respond
to the visible region and show high photocatalytic activity (Das and Daud 2014). The main photo-
catalysts of this type are

• Metal sulfide semiconductors such as CdS and ZnS


• Metal phosphide semiconductors such as GaP and InP
• Others such as AgCl, AgBr, GaAs, and p-Si

Metal sulfides have relatively high CB states and are more appropriate for enhancing solar responses
than metal oxide semiconductors, which are facilitated by the higher VB states consisting of S 3p
orbitals. Especially, CdS with a narrow bandgap of 2.63 eV shows higher catalytic activity for CO2
180 Solar Energy Conversion and Storage: Photochemical Modes

TABLE 9.1
Typical Metal Oxide Semiconductors for Photocatalytic Reduction of CO2 with
Relatively High Yields
Major Product/Yield
Catalyst Bandgap Energy (eV) (μmol/h/g-catalyst) Reference
TiO2 (JRC-TIO) 3.47 CH4/0.03 Anpo et al. (1995)
TiO2 3.0 HCHO/16 Inoue et al. (1979)
ZnO 3.2 HCHO/17.14 Inoue et al. (1979)
W18O49 2.7 CH4/666 ppm/h/g- Xi et al. (2012)
catalyst
HNb3O8 3.66 CH4/3.58 Li et al. (2012)
BiVO4 2.24 C2H5OH/110 Liu et al. (2009)
Montmorillonite-modified 3.07 CH4/441.5 Tahir and Amin (2013)
TiO2
23.2% AgBr/TiO2 2.9 CH4/25.72; Asi et al. (2011)
C2H5OH/15.57
1% CuO/TiO2 3.2 HCOOCH3/1602 Qin et al. (2011)
3% CuO/TiO2 2.88 CH3OH/442.5 Slamet et al. (2005)
5% ZnO/C 3.2 CO/200000 Gokon et al. (2003)
1.7% Cu-0.9% Pt/TiO2 — CH4/33 Zhai et al. (2013)
Pt-KNbO3 3.1 CH4/70 ppm/h/g-catalyst Shi and Zou (2012)
N-InTaO4/Ni-NiO 2.28 CH3OH/165 Tsai et al. (2011)
Sensitized 45% Cu-TiO2 on 3.2 CH3OH/475 Yang et al. (2009)
SBA-15
Ag-BaLa4Ti4O15 — CO/73; HCOOH/2.3 Iizuka et al. (2011)

reduction as compared to other metal sulfides. When it is modified with Bi2S3, the activity can be
improved to some extent (Li et al. 2011). Bi2S3 has a higher photochemical activity and visible light
response than CdS.
The product yields when nonoxide semiconductor photocatalysts are used for CO2 reduction are
still far lower than expected (Table 9.2).
Theoretically, the wide-bandgap semiconductors could be better photocatalysts for CO2 reduc-
tion in aqueous solutions because photogenerated electrons in the bottom of the CB can have suffi-
cient negative redox potential to drive CO2 reduction, while the photogenerated holes in the VB can

TABLE 9.2
Typical Nonoxide Semiconductors for Photocatalytic Reduction of CO2 with a
Relatively High Yield
Major Product/Yield
Catalyst Bandgap Energy (eV) (μmol/h/g-catalyst) Reference
CdS 2.4 HCHO/29; CH3OH/17 Inoue et al. (1979)
SiC 3.0 HCHO/14; CH3OH/76 Inoue et al. (1979)
GaP 2.3 HCHO/14; CH3OH/16 Inoue et al. (1979)
15% Bi2S3-CdS 1.28 CH3OH/120 Li et al. (2011)
p-GaAs — HCOOH/170 μmol Aurian-Blajeni et al. (1983)
RuO2 or Rh1.32Cr0.66O3- — CH3OH/118.5 Liu et al. (2011)
loaded CuxAgyInzZnkSm
Ag@AgX (X = Cl, Br) — CH3OH/45.3; C2H5OH/65.8 Jiang et al. (2010)
Photocatalytic Reduction of Carbon Dioxide 181

be sufficiently energetic to act as acceptors and oxidize water to O2 (Jiang et al. 2010). However, the
wide-bandgap semiconductors cannot use visible light efficiently. Although many nonoxide semi-
conductors can absorb visible light, only a few of them are catalytically active because the energy
levels of neither CBs nor VBs are suitable for CO2 reduction and/or water oxidation. Furthermore,
many of these nonoxide semiconductors suffer from photocorrosion.

9.5  OPERATION CONDITIONS


9.5.1  Pressure Effect
Numerous photocatalysts have been applied for the reduction of CO2. A green chemical approach
is application of a stable photocatalyst, CO2, and water for the production of fuels and chemicals.
In this green approach, the conversion efficiency of CO2 is greatly dependent on the molar ratio
of H2O to CO2. The optimized H2O/CO2 molar ratio is near 5.0 for TiO2 (Anpo et al. 1995). Thus,
how to dissolve more CO2 in H2O becomes an important issue in this process. Increasing CO2
pressure can increase its solubility in H2O. It was found that an increase in CO2 pressure acceler-
ates CO2 reduction in the water, and some products such as ethylene, which are not produced in
ambient CO2 pressure, can be traced under high CO2 pressure. The methanol yield was increased
from 175 to 230  μmol  g-catalyst when CO2 pressure was increased from 110 to 250  kPa, but
decreased to 85  μmol  g-catalyst with further increase of the pressure to 136  kPa (Tseng et al.
2002). Therefore, an optimum CO2 pressure should exist for different photocatalytic reduction
systems.

9.5.2  Reductant Effect


Photoreduction of CO2 competes with H2 formation in the reaction process, when only H2O was
used as the electron donor. In order to solve this problem, it is possible to use other reductants (sac-
rificial electron donors) instead of H2O for CO2 photoreduction. Triethylamine, triethanolamine,
dimethylformamide, acetonitrile, propanol, and dichloromethane with various dielectric constants
are considered as the good sacrificial electron donors (Mikkelsen et al. 2010; Liu et al. 2012). It was
found that the amount of formate increases, while CO decreases with an increase in the dielectric
constant of the reductants. Adding NaOH or other alkaline chemicals in the solutions can dissolve
more CO2 and simultaneously, OH−, in aqueous solutions acts as a strong hole scavenger, enhancing
CO2 conversion. Furthermore, the photoreduction of CO2 can also be performed in the gas phase
using reductants such as CH3OH, CH4, H2, and H2S (Liu et al. 2012).

9.5.3  Temperature Effect


In general, photocatalytic reduction of CO2 is always performed at room temperature. However,
it was found that CO2 reduction can be promoted by increasing the reaction temperature (Richter
et al. 2013). Thermal energy could promote the activity of photocatalyst already activated by
light irradiation. The desorption rate of the products produced on the photocatalyst surface could
be improved by heating. However, CO2 and reductant adsorptions on the photocatalyst surface
should also be controlled by heating. Therefore, an optimum temperature might also exist in a
photoreduction system. Thermal reduction of CO2 will occur if the temperature is too high. In
our previous studies (Guan et al. 2003a, 2003b), photoreduction of CO2 with H2O was performed
under concentrated sunlight using photocatalysts combined with Fe- and Cu/ZnO-based thermal
CO2 reduction catalysts. The reaction temperature reached over 600 K by concentrating the solar
irradiation, and CH3OH and C2H5OH were generated together in the presence of Fe-based CO2
reduction catalyst while only CH3OH was detected in the presence of Cu/ZnO-based thermal CO2
reduction catalyst.
182 Solar Energy Conversion and Storage: Photochemical Modes

9.5.4  Reactor-Type Effect


The physical structure of a photoreactor is very important to ensure photogenerated photos are
effectively collected by the photocatalyst. The design of the irradiation device including mirror,
reflector, window materials, and shape is a key issue for a photoreactor. Before designing a useful
photoreactor, besides making sure that the reactants contacted with the photocatalyst surface well,
two points should be considered:

• How to suspend or support the photocatalysts for utilizing maximum solar spectrum
• How to concentrate sunlight by a reflecting surface to maximize the light harvesting

The most common types of photoreactors for CO2 reduction are slurry reactor, fixed-bed reactor,
bubble-flow reactor, surface-coated reactor, and optical fiber and monolith reactor (Liu et al. 2012;
Tahir and Amin 2013). Among photoreactors, optical fiber and monolith reactors are considered to be
more prominent due to their higher illuminated surface area and efficient utilization of photon energy.

9.6  CONCLUSIONS AND OUTLOOK


Carbon dioxide is a zero-cost carbon source for the synthesis of carbon fuels and value-added chemi-
cals. The catalyst plays a key role in CO2 conversion. Numerous methods and catalysts have been
developed to realize the high efficient conversion of CO2. The photocatalytic reduction process is one
of the best ways to solve the greenhouse gas emission and fuel crisis using the inexhaustible solar
energy source. However, to date, the conversion efficiency and selectivity are still low. The yields of
the main products are below 500 μmol/h/g-catalyst, and the product distributions are also complicated
so that the separation of the products from an aqueous system is poor, which limits the application
of this process. To solve these problems, the following issues should be considered in future studies:

• The mechanisms of photoreduction of CO2, which include multielectron reductions, acti-


vation of two very stable molecules, CO2 and H2O, the link between the semiconductor
structure and the product selectivity, the adsorption/desorption of CO2, the formation of
intermediate products, and the role of adsorbed H2O, should be clarified for the design of
novel photoreduction systems in order to improve the CO2 conversion efficiency and the
selectivity of some target products, such as CH3OH.
• Novel photocatalysts should have the desired bandgap energy for visible light response
and can withstand photocorrosion, long-term stability, and high selectivity. Many wide-
bandgap semiconductors exhibit relatively long-term stability, but only a small amount of
solar energy can be used. Band engineering to improve the visible light absorption ability
of the wide-bandgap semiconductor via nonmetal doping, coupling with narrow-bandgap
semiconductor, photosensitization by inorganic and/or organic dyes, noble metal with sur-
face plasmon band, and quantum dot sensitization have been applied, but the results are
still unsatisfactory. Further study of this strategy is necessary. The other way is to develop
nanocomposite photocatalysts of semiconductor-cocatalyst, semiconductor-carbon mate-
rials such as CNT and graphene, and semiconductor-semiconductor to replace single-
component semiconductors, which have been identified to improve the CO2 conversion
significantly. But the mechanisms of their function are still unclear, and it is necessary to
improve their activity and selectivity.
• One should pay more attention to operating conditions and photoreactor design. Carbon
dioxide pressure, a dissolved amount of CO2 in water, the pH of the solution, the type of
reductant and its amount, the reaction temperature, and the photoreactor type have a great
effect on the product yield as well as the selectivity. The novel photoreduction system
should be simple, convenient, and efficient for the effective utilization of solar energy.
Photocatalytic Reduction of Carbon Dioxide 183

• It is possible to combine photoreduction of CO2 with other reduction processes, such as


thermal reduction of CO2 and electrocatalytic reduction of CO2, in order to reach the maxi-
mum efficiency for the conversion of CO2 to fuels and/or chemicals. Solar energy can
provide photons and thermal energy simultaneously. Concentrated sunlight can also pro-
duce high temperatures for the thermal reduction of CO2. If these two processes can be
combined, total energy efficiency could be improved to a great extent. Furthermore, solar
energy can be transferred to electricity, and the efficiency of the electrocatalytic reduction
of CO2 is generally much higher than that of the photoreduction process. Therefore, devel-
opment of a photoelectrocatalytic combined process to replace the simple photoreduction
process should be more attractive.

REFERENCES
Anpo, M., H. Yamashita, Y. Ichihashi, and S. Ehara. 1995. Photocatalytic reduction of CO2 with H2O on various
titanium oxide catalysts. J. Electroanal. Chem. 396: 21–26.
Asi, M. A., C. He, M. Su, D. Xia, L. Lin, H. Deng, Y. Xiong, R. Qiu, and X. Li. 2011. Photocatalytic reduc-
tion of CO2 to hydrocarbons using AgBr/TiO2 nanocomposites under visible light. Catal. Today. 175:
256–263.
Aurian-Blajeni, B., M. Halmann, and J. Manassen. 1983. Electrochemical measurement on the photoelectro-
chemical reduction of aqueous carbon dioxide on p-Gallium phosphide and p-Gallium arsenide semicon-
ductor electrodes. Solar Energy Mater. 8: 425–440.
Centi, G., and S. Perathoner. 2009. Opportunities and prospects in the chemical recycling of carbon dioxide to
fuels. Catal. Today. 148: 191–205.
Corma, A., and H. Garcia. 2013. Photocatalytic reduction of CO2 for fuel production: Possibilities and chal-
lenges. J. Catal. 308: 168–175.
Das, S., and W. M. A. W. Daud. 2014. A review on advances in photocatalysts towards CO2 conversion. RSC
Adv. 4: 20856–20893.
Dhakshinamoorthy, A., S. Navalon, A. Corma, and H. Garcia. 2012. Photocatalytic CO2 reduction by TiO2 and
related titanium containing solids. Energy Environ. Sci. 5: 9217–9233.
Earth System Research Laboratory. 2014. Global monitoring division. http://www.esrl.noaa.gov/gmd.
Fan, W., Q. Zhang, and Y. Wang. 2013. Semiconductor-based nanocomposites for photocatalytic H2 production
and CO2 conversion. Phy. Chem. Chem. Phy. 15: 2631–2649.
Fujishima, A., X. Zhang, and D. A. Tryk. 2008. TiO2 photocatalysis and related surface phenomena. Surf. Sci.
Rep. 63: 515–582.
Gokon, N., N. Hasegawa, H. Kaneko, H. Aoki, Y. Tamaura, and M. Kitamura. 2003. Photocatalytic effect of
ZnO on carbon gasification with CO2 for high temperature solar thermochemistry. Sol. Energy Mater.
Sol. Cells. 80: 335–341.
Guan, G., T. Kida, T. Harada, M. Isayama, and A. Yoshida. 2003a. Photoreduction of carbon dioxide with water
over K2Ti6O13 photocatalyst combined with Cu/ZnO catalyst under concentrated sunlight. Appl. Catal.
A: Gen. 249: 11–18.
Guan, G., T. Kida, and A. Yoshida. 2003b. Reduction of carbon dioxide with water under concentrated sunlight
using photocatalyst combined with Fe-based catalyst. Appl. Catal. B: Environ. 41: 387–396.
Handoko, A. D., K. Li, and J. Tang. 2013. Recent progress in artificial photosynthesis: CO2 photoreduction to
valuable chemicals in a heterogeneous system. Curr. Opin. Chem. Eng. 2: 200–206.
Hu, B., C. Guild, and S. L. Suib. 2013. Thermal, electrochemical, and photochemical conversion of CO2 to
fuels and value-added products. J. CO2 Utilization 1: 18–27.
Iizuka, K., T. Wato, Y. Miseki, K. Saito, and A. Kudo. 2011. Photocatalytic reduction of carbon dioxide over
Ag cocatalyst-loaded A La4Ti4O15 (A = Ca, Sr, and Ba) using water as a reducing reagent. J. Am. Chem.
Soc. 133: 20863–20868.
Inglis, J. L., B. J. MacLean, M. T. Pryce, and J. G. Vos. 2012. Electrocatalytic pathways towards sustainable
fuel production from water and CO2. Coord. Chem. Rev. 256: 2571–2600.
Inoue, T., A. Fujishima, S. Konishi, and K. Honda. 1979. Photoelectrocatalytic reduction of carbon dioxide in
aqueous suspensions of semiconductor powders. Nature. 277: 637–638.
Izumi, Y. 2013. Recent advances in the photocatalytic conversion of carbon dioxide to fuels with water and/or
hydrogen using solar energy and beyond. Coord. Chem. Rev. 257: 171–186.
184 Solar Energy Conversion and Storage: Photochemical Modes

Jensen, J., M. Mikkelsen, and F. C. Krebs. 2011. Flexible substrates as basis for photocatalytic reduction of
carbon dioxide. Sol. Energy Mater. Solar Cells. 95: 2949–2958.
Jiang, Z., T. Xiao, V. L. Kuznetsov, and P. P. Edwards. 2010. Turning carbon dioxide into fuel. Phil. Trans. R.
Soc. A. 368: 3343–3364.
Kočí, K., L. Obalová, L. Matějová, D. Plachá, Z. Lacny, J. Jirkovsky, and O. Šolcová. 2009. Effect of TiO2
particle size on the photocatalytic reduction of CO2. Appl. Catal. B: Environ. 89: 494–502.
Kondratenko, E. V., G. Mul, J. Baltrusaitis, G. O. Larrazabal, and J. Perez-Ramirez. 2013. Status and perspec-
tives of CO2 conversion into fuels and chemicals by catalytic, photocatalytic and electrocatalytic pro-
cesses. Energy Environ. Sci. 6: 3112–3135.
Li, X., J. Chen, H. Li, J. Li, Y. Xu, Y. Liu, and J. Zhou. 2011. Photoreduction of CO2 to methanol over Bi2S3/
CdS photocatalyst under visible light irradiation. J. Nat. Gas Chem. 20: 413–417.
Li, X., H. Pan, W. Li, and Z. Zhuang. 2012. Photocatalytic reduction of CO2 to methane over HNb3O8 nanibelts.
Appl. Catal. A., 413–414: 103–108.
Liu, G., N. Hoivik, K. Wang, and H. Jakobsen. 2012. Engineering TiO2 nanomaterials for CO2 conversion/solar
fuels. Sol. Energy Mater. Solar Cells. 105: 53–68.
Liu, J., B. Garg, and Y. Ling. 2011. CuxAgyInzZnkSm solid solutions customized with RuO2 or Rh1.32Cr0.66O3
co-catalyst display visible light-driven catalytic activity for CO2 reduction to CH3OH. Green Chem. 13:
2029–2031.
Liu, Y., B. Huang, Y. Dai, X. Zhang, X. Qin, M. Jiang, and M-H. Whangbo. 2009. Selective ethanol formation
from photocatalytic reduction of carbon dioxide in water with BiVO4 photocatalysts. Catal. Commun.
11: 210–213.
Ma, J., N. Sun, X. Zhang, N. Zhao, F. Xiao, W. Wei, and Y. Sun. 2009. A short review of catalysis for CO2
conversion. Catal. Today. 148: 221–231.
Mao, J., K. Li, and T. Peng. 2013. Recent advances in the photocatalytic CO2 reduction over semiconductors.
Catal. Sci. Technol. 3: 2481–2498.
Mikkelsen, M., M. Jorgensen, and F. C. Krebs. 2010. The teraton challenge. A review of fixation and transfor-
mation of carbon dioxide. Energy Environ. Sci. 3: 43–81.
Mori, K., H. Yamashita, and M. Anpo. 2012. Photocatalytic reduction of CO2 with H2O on various titanium
oxide photocatalysts. RSC Adv. 2: 3165–7231.
Morris, A. J., G. J. Meyer, and E. Fujita. 2009. Molecular approaches to the photocatalytic reduction of carbon
dioxide for solar fuels. Acc. Chem. Res. 42: 1983–1994.
Qin, S., F. Xin, Y. Liu, X. Yin, and W. Ma. 2011. Photocatalytic reduction of CO2 in methanol to methyl formate
over CuO-TiO2 composite catalysts. J. Colloid Interface Sci., 356: 257–261.
Richter, R. K. de, T. Ming, and S. Caillol. 2013. Fighting global warming by photocatalytic reduction of CO2
using giant photocatalytic reactors. Renew Sust. Energy Rev. 19: 82–106.
Shi, H., and Z. Zou. 2012. Photophysical and photocatalytic properties of ANbO3 (A = Na, K) photocatalysts.
J. Phys. Chem. Solids. 73: 788–792.
Slamet, H. W. Nasution, E. Purnama, S. Kosela, and J. Gunlazuardi. 2005. Photocatalytic reduction of CO2 on
copper-doped titania catalysts prepared by improved-impregnation method. Catal. Commun. 66: 91–97.
Stechel, E. B., and J. E. Miller. 2013. Re-energizing CO2 to fuels with the sun: Issues of efficiency, scale, and
economics. J. CO2 Utilization. 1: 28–36.
Tahir, M., and N. S. Amin. 2013. Advances in visible light responsive titanium oxide-based photocatalysts for
CO2 conversion to hydrocarbon fuels. Energy Conv. Manage. 76: 194–214.
Tahir, M., and N. S. Amin. 2013. Photocatalytic reduction of carbon dioxide with water vapors over montmoril-
lonite modified TiO2 nancomposites. Appl. Catal. B: Environ. 142–143: 512–522.
Tran, P. D., L. H. Wong, J. Barber, and J. S. C. Loo. 2012. Recent advances in hybrid photocatalysts for solar
fuel production. Energy Environ. Sci. 5: 5902–5918.
Tsai, C.-W., H. M. Chen, R.-S. Liu, K. Asakura, and T.-S. Chan. 2011. Ni@NiO core-shell structure-modified
nitrogen-doped InTaO4 for solar-driven highly efficient CO2 reduction to methanol. J. Phys. Chem. C.
115: 10180–10186.
Tseng, I.-H., J. C. S. Wu, and H.-Y. Chou. 2004. Effects of sol-gel procedures on the photocatalysis of Cu/TiO2
in CO2 photoreduction. J. Catal. 221: 432–440.
Tseng, I.-H., W.-C. Chang, and J. C. S. Wu. 2002. Photoreduction of CO2 using sol-gel derived titania and
titania-supported copper catalysts. Appl. Catal. B: Environ. 37: 37–48.
Windle, C. D., and R. N. Perutz. 2012. Advances in molecular photocatalytic and electrocatalytic CO2 reduc-
tion. Coord. Chem. Rev. 256: 2562–2570.
Wu, J. C. S., H.-M. Lin, and C-L. Lai. 2005. Photo reduction of CO2 to methanol using optical-fiber photoreac-
tor. Appl. Catal. A: Gen. 296: 194–200.
Photocatalytic Reduction of Carbon Dioxide 185

Xi, G., S. Ouyang, P. Li, J. Ye, Q. Ma, N. Su, H. Bai, and C. Wang. 2012. Ultrathin W18O49 nanowires with
diameters below 1  nm: Synthesis, near-infrared absorption, photoluminescence, and photochemical
reduction of carbon dioxide. Angew. Chem. Int. Ed. 51: 2395–2399.
Yang, H.-C., H.-Y. Lin, Y.-S. Chien, J. C.-S. Wu, and H.-H. Wu. 2009. Mesoporous TiO2/SBA-15, and Cu/TiO2/
SBA-15 composite photocatalysts for photoreduction of CO2 to methanol. Catal. Lett. 131: 381–387.
Zhai, Q., S. Xie, W. Fan, Q. Zhang, Y. Wang, W. Deng, and Y. Wang. 2013. Photocatalytic conversion of carbon
dioxide with water into methane: Platinum and copper(I) oxide co-catalysts with core-shell structure.
Angew. Chem. 125: 5888–5891.
10 Artificial Photosynthesis
Neelam Kunwar, Sanyogita Sharma, Surbhi
Benjamin, and Dmitry Polyansky

CONTENTS
10.1 Introduction........................................................................................................................... 187
10.2 Natural Photosynthesis to Artificial Photosynthesis............................................................. 188
10.3 Approaches in Artificial Photosynthesis............................................................................... 190
10.3.1 Homogeneous Artificial Photosynthesis Systems..................................................... 191
10.3.2 Heterogeneous Artificial Photosynthesis Systems.................................................... 195
10.4 Light Absorption.................................................................................................................... 196
10.5 Charge Separation by Molecular Donor–Acceptor Systems................................................. 197
10.6 Light-Driven Catalytic Water Splitting................................................................................. 199
10.7 Catalytic Water Oxidation..................................................................................................... 199
10.7.1 Ruthenium-Based Catalysts.......................................................................................200
10.7.2 Iridium-Based Catalyst..............................................................................................202
10.7.3 Mn-, Co-, and Fe-Based Catalysts............................................................................. 203
10.8 Catalytic Proton Reduction and Hydrogen Evolution............................................................204
10.8.1 Proton Reduction Catalyst by Hydrogenases.............................................................204
10.8.2 Light-Driven Catalysis for Hydrogen Production......................................................206
10.9 Catalysis for CO2 Reduction and Fuel Production................................................................208
References....................................................................................................................................... 213

10.1 INTRODUCTION
Energy is one of the key components sustaining life on Earth, including human existence. The
technological progress of human civilization resulted in a steep increase in anthropogenic energy
consumption, leading to rapid exploration of natural resources of fossil fuels. As on today, approxi-
mately 150,000 TW⋅h of energy is consumed globally per year with over 85% of that still coming
from nonrenewable sources; demand is projected to increase up to 30% in the next 20 years (IEA
2014). The extensive reliance on fossil fuels is associated with several major risk factors, which
highlight the necessity of finding alternative energy sources. One of these factors is the limited
resource base, which when combined with the rapidly increasing demand for energy would result
in the depletion of available natural resources. The uneven distribution of global fossil fuel depos-
its creates another challenge associated with the security of energy supplies, which is greatly
dependent on geopolitical climate. And probably one of the most immediate problems resulting
from the extensive exploration of nonrenewable energy sources is environmental impact, mani-
fested in pollution and global climate change. The latter has drawn significant political attention in
many industrial countries resulting in a wave of new policies, which may ultimately limit the use
of fossil fuels in the future. Altogether, recent trends in the global energy landscape urge extensive
exploration of alternative renewable energy sources (Figure 10.1).
Among many renewable energy sources that are currently being explored, solar energy pres-
ents the most abundant resource with over 5 million terawatt hours (TW⋅h) per year available for

187
188 Solar Energy Conversion and Storage

Environmental Security of
impacts supply

Energy
generation
and use

Growing Policy
demand changes

FIGURE 10.1  Major factors affecting global energy use and production.

practical use. However, one of the major challenges in efficient utilization of solar energy is energy
storage. This is dictated by intermittency of solar flux, day/night cycles, and uneven energy demand,
which significantly restricts the direct use of electricity generated from, for example, photovol-
taic devices. One of the most efficient ways to store energy is in the form of chemical fuels. This
provides the highest energy density (amount of energy stored per total weight of fuel and storage
medium) and allows the use of energy in the existing infrastructure, such as transportation. The
types of fuels that can be produced using the energy of solar light are hydrogen and fuels resulting
from carbon reduction, such as alcohols or methane. In addition to these fuels, which can be used by
a consumer directly, other products derived from the light-driven reduction of CO2, such as carbon
monoxide or formic acid, can be used in further industrial processes, such as syngas feedstock (H2/
CO mixture). The use of hydrogen is very attractive, as it is environmentally clean when produced
using solar energy to split water. Water is the only product that is formed when hydrogen is burned
or consumed in a fuel cell.
The production of chemical fuels using energy of light can be achieved using the artificial pho-
tosynthesis approach. This approach derives its inspiration from natural photosynthesis, where the
energy of sunlight is used to convert water and carbon dioxide to carbohydrates, which are used
later to build the plant’s structure and sustain its functions. Artificial photosynthesis uses the same
functional principles as its natural counterpart; however, components responsible for various func-
tionalities are usually different. In this regard, artificial photosynthesis is a bioinspired approach
(Graetzel 1983, Collings and Critchley 2005), where the important structural elements and reaction
features of natural photosynthesis are used in conceptually simpler systems to achieve the results of
natural photosynthesis. However, the solar energy conversion efficiency of natural photosynthesis is
only a few percentages (0.02%–1%), creating additional challenges in creating artificial photosyn-
thetic systems with efficiencies surpassing those of natural systems. If properly implemented, arti-
ficial photosynthesis should be capable of using solar energy to produce clean fuels at attractively
high efficiencies (Cogdell et al. 2010).

10.2  NATURAL PHOTOSYNTHESIS TO ARTIFICIAL PHOTOSYNTHESIS


In the course of photosynthesis, plants harvest energy from the Sun and use it to split water and
reduce carbon dioxide in order to store this energy in the form of chemical bonds. Photosynthesis
reactions occur in two distinct photosynthetic reaction centers: photosystem I (PS I) and photosys-
tem II (PS II). Light absorption by chlorophyll followed by a series of electron-transfer reactions
Artificial Photosynthesis 189

constitute the initial stage of the light-induced set of reactions. This is followed by oxidation of
water to molecular oxygen which takes place at the oxygen-evolving complex (OEC) of PS II. The
OEC is the natural catalytic center that facilitates the reaction of water oxidation, and it consists of a
cluster of four manganese atoms and one calcium atom linked by oxo-bridges (McEvoy et al. 2005;
McConnell et al. 2010).
In PS II, chlorophyll a (i.e., P680) absorbs photon energy, which excites electrons and transfers
them to plastoquinone (Pq), an electron acceptor (Herek et al. 2002). Meanwhile, the photoexcited
P680 (P680*) generates a driving force for extracting electrons from the water molecule at the
calcium-manganese center, which acts as a water-oxidation catalyst (Umena et al. 2011). Umena
et al. (2011) reported that the OEC is a Mn4CaO5 cluster. Several crystal structures of PS II have
been resolved at various resolutions from 3.8 to 1.9 Å. It was observed that five oxygen atoms served
as oxo-bridges linking five metal atoms, and that at least four water molecules were bound to the
Mn4CaO5 cluster; some of them may, therefore, serve as substrates for dioxygen formation. More
than 1300 water molecules were identified in each photosystem II monomer, out of which some
formed extensive hydrogen-bonding networks that may serve as channels for protons or water or
oxygen molecules.
The electron transport chain, which consists of Pq, cytochrome complex, and plastocyanine (Pc),
carries electrons from PS  II to PS  I (Figure 10.2). The electrons are taken through a series of
uphill and downhill steps to generate energy-rich intermediates in PS I reactions. In PS II, elec-
trons excited by P700 are transferred to ferredoxin (Fd) and used to reduce NADP+ to NADPH
through Fd-NADP reductase. The newly formed NADPH is then consumed to fix CO2 through the
biocatalytic cascade of the Calvin cycle (Hohmann-Marriott and Blankenship 2011) when a series
of light-independent (dark) reactions occur, where the products of the light-induced reactions (ATP
and NADPH) are used to form C–C covalent bonds of carbohydrates, where the carbon source is
CO2, which is obtained from the atmosphere (Balzani et al. 2008).
Natural photosystems give an impetus to the work on artificial photosynthesis. In artificial
photosynthesis, synthetic chromophores or even semiconductor materials can be used instead of
chlorophyll, which can absorb solar light and produce charges with potential enough to, for exam-
ple, drive the oxidation of water to produce O2 and H+. The artificial photosynthetic systems are not
bound to the production of carbohydrates like plants but can be tuned to produce various products
(e.g., H2 through proton reduction) (Sayama et al. 2002), or they can include reduction of CO2 to
carbon-based fuels (methanol, methane, or formate).
Synthetic materials used in artificial photosynthesis have to efficiently realize various functions
of components in natural photosynthesis. Light-absorbing components should efficiently capture

Primary
electron 2e–
acceptor Enzyme
Primary reaction
electron 2e–
acceptor 2e–
ETC

SUN 2e–
2e– NADPH
P700
Photon
ATP
P680 Photon
H2O Photosystem I

½ O2 + 2 H+ Photosystem II

FIGURE 10.2  Electron transfer in the photosystem.


190 Solar Energy Conversion and Storage

and transform photons of light into charges and transport them to catalytic centers, where synthetic
catalysts should efficiently oxidize water to protons and O2 on the oxidative side and reduce protons
to H2 or CO2 to alcohols or methane on the reductive side. As an example of artificial design, an
“artificial leaf” was introduced by Nocera and co-workers. The construction of an artificial leaf
incorporates earth-abundant elements by interfacing a triple-junction, amorphous silicon photovol-
taic with hydrogen and oxygen and evolving catalysts made from a ternary alloy (NiMoZn) and a
cobalt-phosphate cluster (Co-Pi). It captures the structural and functional attributes of the combined
PS II and PS I. The Co-Pi catalyst oxidizes H2O by a proton-coupled electron transfer mechanism,
where oxidation states Co-Pi are increased by four hole equivalents, similar to the S-states pumping
of the Kok cycle of PS II (Nocera 2012).

10.3  APPROACHES IN ARTIFICIAL PHOTOSYNTHESIS


Understanding the biological processes of photosynthesis and utilizing this knowledge to design
some novel synthetic molecular systems to provide good light to chemical energy conversion effi-
ciency is the developing field of artificial photosynthesis (Barber 2009; Benniston and Harriman
2008; Gust et al. 2008). Different approaches have to be based on relationships, which are similar
to those present in a photosynthetic apparatus in living organisms. The main characteristics of
an artificial photosynthetic system must be obtained, taking into account nonvalence interactions
between components of the corresponding model system. The cornerstone of successful design of
functional artificial photosynthesis systems is in the development of efficient, robust, and syntheti-
cally accessible components for light absorption, charge separation, and catalysis, and the provision
of efficient electronic and mass transport between these components. In general, one of the artificial
photosynthetic architectures is envisioned in the form of a photoelectrochemical cell, where reduc-
tive and oxidative half-reactions take place in separate compartments separated by a proton-per-
meable membrane. Electrodes on each side contain catalysts for water oxidation and, for example,
reduction of H+ to hydrogen, respectively. The electrolysis is driven by the energy of solar radiation
absorbed by chromophores or semiconductor materials and converted to electrical charges. The
light-absorbing material can be located either outside the electrolysis cell (PV-to-electrolysis archi-
tecture) or deposited directly on the electrode.
The basic components of artificial photosynthetic systems include (1) light absorption, (2)
excited-state electron transfer, (3) separation of electron transfer–generated oxidative and reductive
equivalents by free energy gradients, (4) electron transfer activation of multielectron catalysts, (5)
proton-electron coupled solar fuel half reactions, and (6) separation and collection of the products
(Figure 10.3).

Homogeneous Heterogeneous
system system

Artificial
photosynthesis

FIGURE 10.3  Systems of artificial photosynthesis.


Artificial Photosynthesis 191

Recently, many promising models have appeared by applying the basic concepts of this technol-
ogy. Two major systems emerging from this are

• Homogeneous artificial photosynthesis systems


• Heterogeneous artificial photosynthesis systems

10.3.1  Homogeneous Artificial Photosynthesis Systems


Homogeneous artificial photosynthesis is a multielectron, multiproton process requiring the use
of catalysts. The design of a homogeneous catalyst is more controllable than that of heterogeneous
systems, and as a result their use can provide better insight into the reaction mechanism. The devel-
opment of homogeneous catalysts is facilitated by access to a wide variety of spectroscopic, kinetic,
and theoretical methods and the ability to relay obtained information to the functioning of the natu-
ral photosynthetic system. Some necessary components of an artificial photosynthesis system are
antenna for light harnessing, a reaction center (RC) for charge separation, oxidation and reduction
catalysts, and membrane to separate the generated products.
Allakhverdiev et al. (2009) examined the main pathways of H2 photoproduction by using pho-
tosynthetic organisms and biomimetic photosynthetic systems. The systematic production of
hydrogen and oxygen is shown in Figure 10.4.
Alstrum-Acevedo et al. (2005) reported light absorption and excited-state electron transfer to
create oxidative and reductive equivalents for driving relevant fuel-forming half reactions such as
the oxidation of water to O2 and its reduction to H2. They prepared an “integrated modular assem-
bly,” with separate components for light absorption, energy transfer, and long-range electron trans-
fer by use of free energy gradients, which are integrated with oxidative and reductive catalysts into
single molecular assemblies or on separate electrodes in photoelectrochemical cells. Derivatives
of porphyrins and metalloporphyrins and metal polypyridyl complexes have been most commonly
used for these assemblies.
Sun et al. (1997) prepared a series of compounds for developing models of PS II in green plants.
In these compounds, a photosensitizer, ruthenium(II) tris(bipyridyl) complex (to mimic the func-
tion of P680 in PS  II), was covalently linked to a manganese(II) ion through different bridging
ligands. The structures of the compounds were characterized by electron paramagnetic resonance
(EPR) measurements and electrospray ionization mass spectrometry (ESI-MS). The interaction
between the ruthenium and manganese moieties within the complex was probed by steady-state and
time-resolved emission measurements. They observed that on exposing the dinuclear complexes to
flash photolysis in the presence of an electron acceptor such as methyl viologen (MV2+), the initial

hν hν

Antenna
4 H+ + O2 2 H+

Catox C D P A C Catred

2 H2O H2

Essential components

FIGURE 10.4  The artificial photosynthetic system for production of oxygen and hydrogen. (P, photosensi-
tizer; A, electron acceptor; D, electron donor; C, electron carrier; Catox, catalyst for oxidation of water; Catred,
catalyst for reduction of H+.) (Adapted from Allakhverdiev, J., Photochem. Photobiol. B, 104, 1–8, 2011. With
permission.)
192 Solar Energy Conversion and Storage

electron transfer from the excited state of Ru(II) in this compound formed Ru(III) and MV+∙. An
intramolecular electron transfer from coordinated Mn(II) to the photogenerated Ru(III) occurred
with a first-order rate constant of 1.8 × 105 s–1, regenerating Ru(II). Thus, a supramolecular system
with a manganese complex has been used as an electron donor to a photooxidized photosensitizer.
Balzani et al. (1997) suggested that the model systems can be developed, which are capable of
mimicking the two fundamental steps of photosynthesis, namely, light harvesting and photoinduced
charge separation.
The theoretical maxima of solar energy conversion efficiencies and productivities in oxygenic
photosynthesis in a variety of photosynthetic organisms, including green microalgae, cyanobac-
teria, and C4 and C3 plants were evaluated by Melis (2009). He also reported that photosynthetic
solar energy conversion efficiency and productivity can be improved by minimizing, or truncat-
ing, the chlorophyll antenna size of the photosystems up to threefold. The generation of truncated
light-harvesting chlorophyll antenna size (tla) strains, in all classes of photosynthetic organisms,
would help to reduce excess absorption of sunlight and the ensuing wasteful dissipation of excita-
tion energy. It may also maximize solar-to-product energy conversion efficiency and photosynthetic
productivity in high-density mass cultivations. Such a model may find application in the commercial
exploitation of microalgae and plants for the generation of biomass, biofuels, and chemical feed-
stocks, as well as in the field of nutraceuticals and pharmaceuticals.
Absorption of sunlight by a chromophore is followed by the transfer of excitation energy to a
reaction center (C). However, light absorption by a single molecule of a chromophore is limited and
requires either the use of antenna arrays or multilayer structures for harnessing the light efficiently.
Excitation of a chromophore molecule is then followed by photoinduced electron transfer to an
electron acceptor (A), and a secondary thermal electron transfer from a donor component (D) to an
oxidized chromophore (C*) (Figure 10.5). If the antenna array and reaction center (C) are not prop-
erly coupled in the dimensions of time, energy, and space, homogeneous artificial photosynthetic
systems may not function properly.
Ort et al. (2011) discussed the theoretical and physiological factors that need to be considered in
optimizing chlorophyll content for improved photosynthetic efficiency and maximum carbon gain
by crops and microalgae. It was concluded that on optimizing the size of antenna (by reducing its
size), the photosynthetic efficacy was also maximized. Multiporphyrin arrays, polypyridine com-
plexes, and polymetallic complexes were among the molecules that have been widely investigated as
antenna molecules. Balzani et al. (2008) reviewed a brief description of the mechanism at the basis
of natural photosynthesis and its relevance to the field of photochemical conversion of solar energy.
Solar energy, whether converted to wind, rain, biomass, or fossil fuels, was found to be the primary
energy source for human-engineered energy transduction systems.
Hambourger et al. (2009) reviewed matches between some biological and technological energy
systems. Herrero et al. (2008) described various aspects of the oxidizing enzyme or designing arti-
ficial catalysts, which have the ability to oxidize water with a low overpotential and could also
greatly improve the efficiency of water electrolysis and photolysis. They provided an overview of
the production of a biomimetic photocatalyst for water oxidation by focusing on the essential cofac-
tors involved in the light-driven oxidation of water.

D A

FIGURE 10.5  Excitation of chromophore.


Artificial Photosynthesis 193

An understanding of natural photosynthesis at the molecular level has been supported further
by the creation of artificial photosynthetic model systems, such as donor−acceptor assemblies. The
effort has been directed toward the development of artificial systems composed of molecular and
supramolecular architectures. In molecular light-harvesting systems, the energy of many photons is
collected and transferred to a final acceptor, resulting in charge separation. Light energy is used to
move an electron from donor to acceptor, creating this charge-separated state. The resulting oxidiz-
ing and reducing equivalents are moved away from the site of charge separation to catalytic sites
through subsequent electron transfer steps, where they are used to oxidize water to oxygen and carry
out reductive interaction that can form fuels in oxygenic photosynthesis. Photoinduced electron
transfer is also characteristic of these artificial reaction centers. Such systems have been developed
to mimic the antenna effect.
Dyad and triad models based on the association of a photosensitizer to an electron donor, an
electron acceptor, or both components, have also been designed to reproduce light-induced spatial
charge separation. These samples show that transformation of light energy into chemical energy in
the form of a charge-separated state can be achieved. However, the next step toward water decom-
position would require the coupling of the relevant reduction and oxidation catalysts to a triad. The
basic requirements for such a particular working photocatalytic system are

• Directionality of the electron transfer process for each individual charge separation step
• Tuning of the electronic coupling between the donor/acceptor and the photosensitizer in
order to maximize electron transfer efficiency
• Long-lived charge-separated state
• Catalysts (donor/acceptor) capable of storing multiple oxidizing/reducing equivalents

Gust et al. (2001) reported that artificial light-harvesting antenna can be synthesized and linked
to artificial reaction centers that convert excitation energy to chemical potential in the form of
long-lived charge separation. These artificial reaction centers can form the basis for molecular-
level optoelectronic devices. Fukuzumi (2008) presented a major strategy to reach long-lived charge
separation states, which is an important requirement in view of the application of photogenerated
electrons in catalysis.
Aratani et al. (2009) reviewed the synthesis of covalently and noncovalently linked discrete cyclic
porphyrin arrays as models of the photosynthetic light-harvesting antenna complexes. They also pre-
pared a series of extremely long but discrete meso-meso-linked porphyrin arrays and covalently linked
large porphyrin rings on the basis of the silver(I)-promoted oxidative coupling strategy. The photo-
physical properties of these molecules were examined using steady-state absorption, fluorescence,
fluorescence lifetime, fluorescence anisotropy decay, and transient absorption measurements. The
exciton-exciton annihilation time and the polarization anisotropy rise time within these structures
were described in terms of the Forster-type incoherent energy hopping model.
Meso-pyridine-appended zinc(II) porphyrins and their meso-meso-linked dimers spontaneously
assemble in noncoordinating solvents such as CHCl3 to form tetrameric porphyrin squares and
porphyrin boxes, respectively. The rigorous homochiral self-sorting process and efficient excitation
energy transfer (EET) have been demonstrated along these cyclic porphyrin arrays. The meso-
cinchomeronimide appended zinc(II) porphyrin forms a cyclic trimer. It has also been shown that
the corresponding meso-meso-linked diporphyrins undergo high-fidelity self-sorting assembling to
form discrete cyclic trimer, tetramer, and pentamer with large association constants through per-
fect discrimination of enantiomeric and conformational differences of the meso-cinchomeronimide
substituents (Naoki et al. 2009).
Wenger (2009) reported the long-range electron transfer with synthetic systems using d6 or d8
metal complexes as photosensitizers, which includes charge transfer over distances greater than
10  Å in both covalent and noncovalent donor–bridge–acceptor systems. The coordination com-
pounds studied contained metals such as Ru(II), Os(II), Re(I), Ir(III), Ir(I), and Pt(II). Albinsson
194 Solar Energy Conversion and Storage

and Martensson (2008) reported that the property of a donor-bridge-acceptor (D-B-A) system is
not a simple linear combination of properties of the individual components, but it depends on the
specific building blocks and the style in which they are assembled. An important example is the
ability of the bridge to support the intended transfer process. The mediation of the transfer is char-
acterized by an attenuation factor (β) often viewed as a bridge-specific constant, but it also depends
on the donor and the acceptor (i.e., the same bridge can either be poorly or strongly conducting
depending on the nature of the donor and the acceptor). An account of the experimental exploration
of the attenuation factor β in a series of bis(porphyrin) systems covalently linked by bridges of the
oligo(phenyleneethynylene) (OPE) type has been given.
Attenuation factors for electron transfer (ET) as well as for both singlet and triplet excitation
energy transfer (EET) are deliberated. A report is also given on the dependence of the transfer
efficiency on the energy gap between the donor and bridge states relevant for the specific transfer
process. The experimental variation of β with varying donor and acceptor components is shown for
a range of conjugated bridges by representative examples. The theoretical rationalization for the
observed variation is briefly discussed. Based on the Gamow tunneling model, the observed varia-
tions in β-values with varying donors and acceptors for the same bridges are simulated, and the
observed energy gap dependence is modeled by Bo and Martensson (2008).
Harriman and Sauvage (1996) described a new approach for assembling the photosynthetic mod-
els with porphyrin-containing modules assembled around transition metals. Straight et al. (2008)
prepared a molecular pentad consisting of two light-gathering antennas, a porphyrin electron donor,
a fullerene electron acceptor, and a photochromic control moiety. This molecular system mimics
the nonphotochemical quenching mechanism for photoprotection found in plants where the light
conversion efficiency is altered in response to the incident photon flux. Specifically, this molecule
assembly achieves photoinduced electron transfer with a quantum yield of 82% when the light
intensity is low. As the light intensity was increased, the photochrome photoisomerizes, leading
to the quenching of the porphyrin-excited state and reducing the quantum yield to as low as 27%
(Figure 10.6).
The transition-metal coordination compounds are often used as photocatalysts for the reduction
of carbon dioxide due to their ability to absorb a significant portion of the solar spectrum and pro-
mote activation of small molecules as reported by Morris et al. (2009). The review discussed four
classes of transition-metal catalysts: (1) metal tetraaza-macrocyclic compounds; (2) supramolecular
complexes; (3) metalloporphyrins and related metallomacrocycles; and (4) Re(CO)3(bpy)X-based
compounds (where bpy is 2,2′-bipyridine). Carbon monoxide and formate are the primary CO2
reduction products, in addition to the formation of hydrogen and bicarbonate/carbonate.
Yin et al. (2010) reported the complex [Co4(H2O)2(PW9O34)2]10–, which includes a Co4O4 core
and is stabilized by oxidatively resistant polytungstate ligands. It was described as a molecular fully
inorganic water oxidation catalyst, which is hydrolytically and oxidatively stable and self-assembles
in water from salts of earth-abundant elements (Co, W, and P). Substantial catalytic turnover

D P

P A

D P A

FIGURE 10.6  Dyad and triad models mimicking the charge separation state.
Artificial Photosynthesis 195

frequencies (TOFs) for O2 production greater than or equal to 5 s–1 at pH = 8 were observed with
[Ru(bpy)3]3+ (bpy is 2,2′-bipyridine) as the photogenerated oxidant. The sensitivity of the catalytic
rate to the pH change reflects the pH dependence of the rate-limiting step in the overall four-electron
H2O oxidation. Extensive spectroscopic, electrochemical, and inhibition studies indicated that
[Co4(H2O)2(PW9O34)2]10− is a stable complex under catalytic turnover conditions.

10.3.2  Heterogeneous Artificial Photosynthesis Systems


The use of heterogeneous catalysts is ultimately more attractive for commercial-scale solar fuel
production due to their higher stability toward oxidative degradation and more facile and eco-
nomical fabrication. Moreover, heterogeneous catalysts can be more easily integrated into devices
that are capable of combining processes of the water oxidation and proton reduction, making the
complete artificial system functional. The term artificial photosynthesis, was initially applied to
molecular systems, but now it is also applied for water-splitting and fuel-forming reactions at the
semiconductor–electrolyte interface. Upon irradiation of a semiconductor-based photocatalyst with
light having energy greater than the semiconductor’s bandgap, electrons are promoted to the con-
duction band (CB), while holes are left in the valance band (VB). However, hole–electron pairs are
still bound by Coulombic forces forming an exciton. These light-generated electron–hole pairs can
be considered as localized versions of reductive and oxidative equivalents. Once formed, these elec-
tron–hole pairs are separated by internal electric fields and directed to spatially separate catalytic
sites on the semiconductor surface. The photogenerated electrons and holes may undergo

• Recombination in bulk
• Recombination at the surface
• Reduction of suitable electron acceptor or oxidation of an electron donor

Electron–hole recombination is promoted by defects in the semiconductor materials and decreases


the efficiency of photocatalytic activity. If an electron donor molecule (D) is present at the sur-
face, then the photogenerated hole can react with this molecule to generate oxidizing products, D+.
Similarly, if an electron acceptor molecule (A) is present at the surface, then the photogenerated
electrons can react with them to generate a reduced product, A− (Prashant et al. 2003):


A + D ⎯Semiconductor
⎯⎯⎯⎯ → A− + D+ (10.1)

In order to increase the light collection efficiency of wide-band semiconductors, their surfaces can
be sensitized with molecular chromophores or dyes. In addition to extension of light-absorbing
properties to the visible part of the solar spectrum, which increases the amount of photons collected
by the absorber, dye molecules improve the efficiency of charge separation, due to ultrafast electron
injection from the excited states of chromophores into the valence bands of semiconductor materi-
als. Such dye/semiconductor interfaces make the basis of a dye-sensitized solar cell (DSSC). DSSC
can also be considered as an example of a heterogeneous artificial photosynthesis system, which
mimics the electron transfer processes in natural photosynthesis. While light-activated electron
transfer processes in the natural photosystem produce charge gradient in the form of proton-motive
force (pmf), the action of DSSC results in the production of electromotive force (emf), or in other
words, electricity.
Abe et al. (2000) developed a stable dye-sensitized photocatalyst system. They reported that eosin
Y attached to Pt-TiO2 (EY-TiO2) exhibited steady H2 production from aqueous triethanolamine solu-
tion (TEOA aq.) under visible light irradiation for an extended time. The turnover number (TON)
of the light-driven hydrogen production reached more than 10,000, and the quantum yield of the
EY-TiO2 at 520 nm was determined to be about 10%.
196 Solar Energy Conversion and Storage

However, regardless of the mechanism of the generation of positive and negative charges, all
current heterogeneous photosynthetic systems can be described in terms of the photoelectrochemi-
cal cell (PEC) concept, where electrons are transferred from a photoanode by oxidizing water to
a cathode to reduce protons or CO2 in the fuel-forming reactions. Bak et al. (2002) discussed the
materials-related issues in the development of high-efficiency PECs. The property requirements for
photoelectrodes, in terms of light absorption and electrochemical properties and their impact on the
performance of PECs, were defined. Different types of PECs were overviewed, and the impacts of
the PEC structure and materials selection on the conversion efficiency of solar energy were consid-
ered. They also reported that the performance of PECs is to be considered in terms of

• Excitation of electron–hole pair in photoelectrodes


• Charge separation in photoelectrodes
• Electrode processes and related charge transfer within PECs
• Generation of the PEC voltage required for water oxidation and reductive reactions

10.4  LIGHT ABSORPTION


Absorption of solar light is the first step in the sequence of reactions leading to the successful
operation of artificial as well as natural photosystems. Nature has highly complex light-harvesting
systems to utilize a substantial part of the solar spectrum by using chlorophyll dyes. The goal of
artificial systems is to incorporate enhanced light-harvesting capacity and efficiency to mimic natu-
ral systems. In an artificial system, sufficient light harvesting can be achieved by using a single RC
chromophore or by the excitation of a light-absorbing antenna array followed by the energy transfer
sensitization of a reaction center (RC) (Alstrum-Acevedo et al. 2005). Antenna systems are capable
of collecting light and transferring energy in an efficient and direct way. A large number of antenna
chromophores surrounding the RC absorb incident photons if a RC is coupled to an antenna array.
The resulting excited states transfer electronic energy to the RC before undergoing radiative or
nonradiative deactivation. Highly branched tree-like dendrimers have been widely employed as
antenna systems. Melis (2009) reported that the incorporation of antenna chromophores in arti-
ficial systems may not be beneficial, if the antenna array and the RC are not properly coupled in
the dimensions of time, energy, and space. It has been shown that the photosynthetic solar energy
conversion efficiency in natural systems will improve by reducing the antenna size.
Among various light-harvesting chromophores, porphyrins and phthalocyanines are closely
related to chlorophyll derivatives. Röger et al. (2006) prepared a bichromophoric assembly consist-
ing of blue naphthalene bisimide (NBI) dye at the periphery of aggregated zinc chlorines. Metal
coordination compounds, exhibiting metal-to-ligand charge-transfer (MLCT) transitions at rela-
tively low energy, have been widely employed as photosensitizers (Huynh et al. 2005) (Figure 10.7).
It was reported that unlike chlorophyll a, the chromophore can be easily functionalized and incor-
porated into a wide variety of biomimetic electron donor–acceptor systems.
Meyer (1990) reported that polypyridyl complexes of Ru(II), Os(II), and Re(1) absorb visible
light through MLCT transitions. They provided insight into the reactivity of MLCT excited states
through the intramolecular electron and energy transfer processes. Lukas et al. (2002) prepared
a green chromophore, 1,7-bis(pyrrolidin-1-yl)-3,4:9,10-perylene-bis(dicarboximide) (5PDI), that
exhibits photophysical and redox properties similar to those of chlorophyll a (Chl a), but unlike Chl
a, it can be easily functionalized and incorporated into a wide variety of biomimetic electron donor–
acceptor systems. The N,N′-dicyclohexyl derivative (5PDI) absorbs strongly (ɛ = 46,000 M L cm–1)
at 686 nm in toluene and fluoresces at 721 nm with a 35% quantum yield. Additionally, 5PDI can
be oxidized as well as reduced in CH2Cl2 at 0.57 V and –0.76 V versus SCE (saturated calomel
electrode), respectively, making it a facile electron donor or acceptor. Rodlike covalent electron
donor–acceptor molecules 5PDI-PI, 5PDI-NI, and 5PDI-PDI were prepared by linking the imide
Artificial Photosynthesis 197

π∗ (bpy)


2+
450 nm
N
N N dxydx2dy2
Ru
N N
N
π (bpy)

Ground state Excited state


dπ (Ru)6 dπ (Ru)5π∗ (bpy)1

FIGURE 10.7  Energy-level diagram showing the ground and lowest MLCT excited state(s) of [Ru(bpy)3]2+.
(Adapted from Huynh, M. H. V., D. M. Dattelbaum, and T. J. Meyer, Coord. Chem. Rev., 249, 457–483, 2005.
With permission.)

group of the 5PDI donor to pyromellitimide (PI), 1,8:4,5-naphthalene-bis(dicarboximide) (NI),


and 1,7-bis(3,5-di-tert-butylphenoxy)-3,4:9,10-perylene-bis(dicarboximide) (PDI) acceptors via an
N-N bond. The formation and decay of their excited and radical ion pair states were monitored
directly by transient absorption spectroscopy following femtosecond laser excitation of 5PDI,
5PDI-PI, 5PDI-NI, and 5PDI-PDI in toluene and 2-methyltetrahydrofuran. The supramolecular
light-harvesting arrays can be constructed by self-assembling chromophore building blocks in solu-
tion and on surfaces (Ahrens et al. 2004; Kelley et al. 2008).

10.5  CHARGE SEPARATION BY MOLECULAR DONOR–ACCEPTOR SYSTEMS


Charge separation processes play an essential role in the functionality of natural and artificial pho-
tosystems. Dinner and Babcock (1996) and Danielsson et al. (2006) reported the electron transfer
sequence of light-driven water oxidation in PS II:

• P680 is excited by absorption of photons or energy transfer from light-harvesting antenna


chlorophylls, and forms P680*, which transfers an electron to PS I via pheophytin (Phe),
quinones A and B (QA and QB), resulting in a highly oxidizing P680 •+.
• P680 •+ oxidizes tyrosine Z (Tyrz) to a milder Tyrz radical species.
• Tyrz radical extracts an electron from the nearby OEC.
• Water oxidation occurs when four electrons are transferred.

Following the initial event of light absorption, the energy of photons is stored in metastable excited
states, where the photoexcited electron is promoted from the highest occupied molecular orbital
(HOMO) of the chromophore to a higher-energy molecular orbital of the same molecule. This pro-
duces a tremendous energy gradient within the photoexcited molecule, which forces an excited
electron to return to its original state and release the excess energy in the form of heat (internal
conversion) or light (luminescence). Photoexcited molecules can become better electron donors or
acceptors, due to the charge separation character of the excited states. In order to avoid detrimental
recombination processes and maximize the generation of charges, the charge separation character
of the excited state should be enhanced, which can be achieved, for example, by incorporating
donor–acceptor (D–A) assemblies in the light-absorbing units. These donor–acceptor assemblies
facilitate the electron transfer, charge separation, and storage of redox equivalents for multielec-
tron reactions by providing coupling between the donor or acceptor and the chromophore. Some
of the donor–acceptor assemblies are designed as molecular dyads, such as the porphyrin qui-
none (P−Q) system. In the P−Q system, the covalently linked porphyrin and quinone function as
198 Solar Energy Conversion and Storage

the chromophore/electron donor and electron acceptor, respectively. Gust and Moore (1989) have
described features of some dyad models, which mimic the photosynthetic electron transfer and
singlet or triplet energy transfer. They also provided insight into how this type of dyad can be used
for elucidating basic photochemical principles of artificial photosynthetic systems. Some limitations
of these types of dyads in the area of temporal stabilization of electronic charge separation were
also presented. This limitation inspired the development of multicomponent molecular devices with
much more complex structures.
The more complex D–A architectures have also been explored. Moore et al. (1984) reported a
molecular triad constructed by using a porphyrin (P) photosensitizer linked to a carotenoid (Car)
electron donor on one side and a quinone (Q) electron acceptor on the other side. In this Car−
P−Q triad, photoexcitation of the porphyrin yields the first excited singlet state, Car−1P*−Q. This
excited state decays via a sequential, two-step, electron transfer process, leading to the formation
of a Car•+−P−Q•− charge-separated state with a lifetime on the microsecond (μs) timescale. The
charge recombination is significantly slowed because of the interposition of a neutral porphyrin
between the widely separated ions. This triad approach mimics the strategy used in natural RCs,
where multistep electron transfers occur through a series of donors and acceptors. The ability to
optimize charge separation over charge recombination in any system, thereby creating long-lived
charge-separated states, is essential for the development of efficient artificial photosynthetic sys-
tems. The electron transfer rate is a function of donor–acceptor orientation, the solvent, the inter-
vening linkage or other medium, and the temperature in donor–acceptor assemblies.
Wiberg et al. (2007) observed that the donor–acceptor distances and donor bridge energy gaps
influence the rates of charge separation and charge recombination differently in a donor-bridge-
acceptor model system. It was reported that the exponential distance dependence is slightly
higher for charge recombination in comparison with that for charge separation. They also have
shown that the effect of the tunneling barrier height is different for charge separation and charge
recombination, and that this difference is highly dependent on the electron acceptor. Sun et al.
(2001) studied some supramolecular donor–acceptor model systems, which are capable of mim-
icking the light reactions on the donor side of photosystem II (PS  II). They also reported that
manganese complexes and tyrosine play the role of electron donors in the model system, similar
to the manganese cluster of OEC and tyrosine in PS II. The donors have been covalently linked
to a photosensitizer, a ruthenium(II) tris-bipyridyl complex, that plays the role of the P680 chloro-
phylls in PS II. It was observed that the incorporation of an intervening redox active link (such as
tyrosine) might be crucial to the multistep electron transfer from the Mn cluster to the Ru center,
as in PS II. The light-harvesting [Ru(bpy)3]2+ moiety was used as the replacement for P680; Mn
complexes, such as a Mn(II,II) dimer were incorporated in the assemblies to mimic the OEC
(Abrahamsson et al. 2002).
Light-induced accumulative electron transfer from the Mn(II,II) dimer to the photooxidized Ru
center resulted in the formation of a Mn(III,IV) complex (Hammarström and Styring 2008). Huang
et al. (2002) reported that the water-oxidizing complex in photosystem II, a dimeric manganese(II,II)
complex, was linked to a ruthenium(II)tris-bipyridine (RuII(bpy)3) complex via a substituted l-tyro-
sine, to form the trinuclear complex. The photolysis of the dimeric complex and RuII(bpy)3 in aque-
ous solution, in the presence of an electron acceptor, resulted in the stepwise extraction of three
electrons by RuIII(bpy)3 from the Mn2II,II dimer, which then attained the Mn2III,IV oxidation state.
It indicates that oxidation from the Mn2II,II state proceeds stepwise via intermediate formation of
Mn2II,III and Mn2III,III. In the presence of water, cyclic voltammetry showed an additional anodic peak
beyond Mn2II,III/III,III oxidation, which was significantly lower than in neat acetonitrile. Assuming
that this peak is due to oxidation to Mn2III,IV suggests that water is essential for the formation of the
Mn2III,IVoxidation state. A trinuclear complex is a structural mimic of the water-oxidizing complex,
where it links a Mn complex via a tyrosine to a highly oxidizing photosensitizer. This complex also
mimics mechanistic aspects of photosystem II, where the electron transfer to the photosensitizer is
fast and results in several electron extractions from the Mn moiety.
Artificial Photosynthesis 199

10.6  LIGHT-DRIVEN CATALYTIC WATER SPLITTING


The overall reaction of the water splitting represents the transformation of two water molecules into
one oxygen molecule and two hydrogen molecules. This reaction is endoergic and requires energy
input, which can be provided by the energy of solar light:

2H 2O = O 2 + 2H 2 ΔG° = 474 kJ mol−1 (10.2)

The reaction of water splitting is typically separated into two of its half-reactions: catalytic water
oxidation and catalytic proton reduction. Water oxidation is the most energy demanding, as com-
pared with the latter. The difficulty of water oxidation is due to the complexity of the reaction, which
involves multiple proton-coupled electron transfer (PCET) processes and the O–O bond formation:

2H 2O = O 2 + 4e− + 4H + ΔG° = 474 kJ mol−1 (10.3)

2e− + 2H + = H 2 ΔG° = 0 kJ mol−1 (10.4)

In electrochemical terms, the equilibrium potential required for achieving water oxidation is 1.23 V/
NHE at pH = 0 and falls 59 mV per every unit of pH increase. In practice, a potential in excess
of equilibrium potential is required to overcome the formation of high-energy intermediates and
achieve reasonable reaction rates. As a result, a water oxidation catalyst (WOC) is required to mini-
mize the overpotential and increase the rate of water oxidation. A qualified WOC has to fulfill the
following criteria for large-scale application:

• Long-term durability
• Low overpotential
• High activity
• Low cost
• Low toxicity

10.7  CATALYTIC WATER OXIDATION

2H 2O ⎯Photosensitizer
hν,WOC
⎯⎯⎯⎯ → O 2 + 4e− + 4H+ (10.5)

This reaction is an essential part of water splitting. The net half-reaction is water oxidation to
produce electrons and protons, with oxygen being a “by-product.” A simple prototype architec-
ture can be designed to study light-driven water oxidation by combining a water oxidation catalyst
with a light-absorbing chromophore in a single assembly. A series of oligoproline-based light-
harvesting chromophore water oxidation catalyst assemblies were presented by Ryan et al. (2014).
This approach combines solid-phase peptide synthesis and also includes copper(I)-catalyzed azide-
alkyne cycloaddition (CuAAC) as an orthogonal approach to install the chromophore and assemble
the water oxidation catalyst. Alibabaei et al. (2013) presented a strategy for solar water splitting
based on a dye-sensitized photoelectrolysis cell. It uses a derivatized, core–shell-nanostructured
photoanode with the high surface area conductive indium tin oxide or antimony tin oxide core
coated with a thin outer TiO2 shell formed by atomic layer deposition. A chromophore-catalyst
assembly [(PO3H2)2bpy)2Ru(4-Mebpy-4-bimpy) Rub(tpy)(OH2)]4+, which combines a light absorber
and water oxidation catalyst in a single molecule, was attached to the TiO2 shell. Visible light pho-
tolysis of the resulting core–shell assembly containing Pt cathode resulted in water splitting into
200 Solar Energy Conversion and Storage

hydrogen and oxygen. The strong electron-donating oxo and carboxylate ligands in the OEC play an
essential role in stabilizing and lowering redox potentials of the OEC. This was an inspiration for
designing a few families of synthetic WOCs based on transition metals and strong electron-donating
ligands. A variety of homogeneous water oxidation catalysts have been developed which contain
transition metals including Ru, Mn, Ir, Co, and Fe.

10.7.1 Ruthenium-Based Catalysts
Gersten et al. (1982) reported the molecular WOC cis,cis-[Ru(bpy)2(H2O)]2(µ-O)]4+ “blue dimer”
with a TON of about 13 and a TOF of 0.004 s−1. This was the first report of molecular water oxida-
tion catalysts based on transition-metal complex. Thirty years of development on Ru WOCs has led
to a dramatic improvement in the catalytic efficiency. Now, Ru-based WOCs are the most exten-
sively studied systems, not only for the purpose of the mechanistic investigation of the water oxida-
tion reaction but also for providing new insights into the structure-activity relationship.
Romain et al. (2009) reported that Ru–aqua complexes are capable of reaching high oxidation
states as a result of the sequential or simultaneous loss of protons and electrons (Figure 10.8).
A solvent water molecule may or may not participate in the formation of the O−O bond.
Accordingly, the two main pathways are

1. Solvent water nucleophilic attack (WNA) on the Ru=O group.


2. Interaction of two M−O units (I2M). Most of the complexes described belong to the WNA
class, including a variety of mononuclear and polynuclear complexes containing one or
several Ru−O units.

A common feature of these complexes is the generation of formal oxidation states as high as Ru(V)
and Ru(VI), which render the oxygen atom of the Ru−O group highly electrophilic. On the other
hand, only one symmetric dinuclear complex that undergoes an intramolecular O−O bond forma-
tion step has been described for the I2M class. It has a formal oxidation state of Ru(IV). It was also
observed that Ru−OH2 complexes contain redox active ligands (such as the chelating quinone).
These ligands are capable of undergoing reversible redox processes, and thus generate a complex
but fascinating electron transfer process between the metal and the ligand.
A general reactivity toward water oxidation in a class of Ru polypyridine complexes was
investigated by tuning the properties of molecular catalysts by systematic synthetic variations of
ligand manifolds (Concepcion et al. 2009). These molecules catalyze water oxidation driven either

H
H
Ru H
H
H Ru
WNA H 12M
H H
H H
H
Ru H Ru H
H
H Ru Ru
H
H H H
H
H

FIGURE 10.8  Ruthenium aqua complex. (Adapted from Romain, S., L. Vigara, and A. Llobet, Acc. Chem.
Res., 42, 1944–1953, 2009. With permission.)
Artificial Photosynthesis 201

electrochemically or by Ce(IV). The first two in the series were [Ru(tpy)(bpm)(OH2)]2+ and [Ru(tpy)
(bpz)(OH2)]2+ (bpm is 2,2′-bipyrimidine; tpy is 2,2′:6′,2″-terpyridine), which undergo hundreds of
turnovers without decomposition, with Ce(IV) as an oxidant. Detailed mechanistic studies and
DFT calculations have revealed a stepwise mechanism: initial 2e−/2 H+ oxidation, to RuIV ═O2+,
1e− oxidation to RuV ═O3+, nucleophilic H2O attack resulting in RuIII−OOH2+, further oxidation to
RuIV(O2)2+, and, finally, oxygen loss, which is in competition with further oxidation of [RuIV(O2)]2+
to [RuV(O2)]3+, which loses O2 more rapidly.
An extended family of 10−15 catalysts based on Mebimpy (Mebimpy is 2,6-bis(1-methyl-
benzimidazol-2-yl)pyridine), tpy, and heterocyclic carbene ligands appears to share a common
mechanism. The osmium complex [Os(tpy)(bpy)(OH2)]2+ also functions as a water oxidation cata-
lyst. Mechanistic experiments have revealed additional pathways for water oxidation: one involving
Cl− assisted catalysis, and another rate enhancement of O–O bond formation by concerted atom
proton transfer (APT).
Sens et al. (2004) synthesized three new dinuclear ruthenium complexes. These are [Ru2II(bpp)
(trpy)2(μ-L)]2+ (L = Cl or AcO) and [Ru2II(bpp)(trpy)2(H2O)2]3+. These three complexes have been
characterized through the usual spectroscopic and electrochemical techniques. In aqueous acidic
solution, the acetato bridge of the second complex is replaced by aqua ligands, generating the
bis(aqua) complex, which is oxidized to its RuIVRuIV state. It has been shown to catalytically oxi-
dize water to a molecular oxygen. The measured pseudo-first-order rate constant for the O2 evolu-
tion process is 1.4 × 10 –2  s–1, which is more than three times larger than that reported earlier for
Ru-O-Ru–type catalysts. This new water-splitting catalyst also has some improved stability.
Wada et al. (2000) synthesized the bridging ligand 1,8-bis(2,2′:6′,2″-terpyridyl) anthra-
cene (btpyan). The resulting dinuclear complex [Ru2(II)(OH)2(3,6-(t)Bu2sq)2(btpyan)]0 under-
goes ligand-localized oxidation at E1/2  =  +0.40  V (vs. Ag/AgCl) to give [Ru2(II)(OH)2(3,6-(t)
Bu2qui)2(btpyan)]2+ in MeOH solution. Furthermore, metal-localized oxidation of [Ru2(II)(OH)2(3,
6-(t)Bu2(qui)2(btpyan)]2+ at E-p = + 1.2 V in CF3CH2OH/ether or water gives [Ru2(III)(OH)2(3,6-(t)
Bu2(qui)2(btpyan)]4+, which catalyzes water oxidation. Controlled potential electrolysis of [Ru2(II)
(OH)2(3,6-(t)Bu2qui)2(btpyan)]2+ (SbF6)2 at +1.70 V in the presence of H2O in CF3CH2OH evolves
dioxygen with a current efficiency of 91% (21 turnovers). The turnover number of O2 evolu-
tion increases to 33,500 when the electrolysis is conducted in water (pH 4.0) by using a [Ru2(II)
(OH)2(3,6-(t)Bu2qui)2(btpyan)]2+ (SbF6)2-modified ITO electrode (Wada et. al. 2001).
A dinuclear ruthenium complex has been synthesized by Xu et al. (2010) (Figure 10.9).
It was employed to catalyze the homogeneous water oxidation. An exceptionally high TON was
observed in the cases of chemical (CeIV as the oxidant) and light-driven ([Ru(bpy)3]2+-type photo-
sensitizers) water oxidation. Cyclometalated dimeric ruthenium(II) complexes [Ru2(cppd)(pic)6]+

Red. CeIII

RuII
O2 + 4H+

RuIII
Ox.

CeIV
2 H2O

FIGURE 10.9  Homogeneous water oxidation. (Adapted from Xu, Y., A. Fischer, L. Duan, L. Tong, E.
Gabrielsson, B. Åkermark, and L. Sun, Angew. Chem. Int. Ed., 49, 8934–8937, 2010. With permission.)
202 Solar Energy Conversion and Storage

N
SnBu3 [Ru(DMSO)4Cl2]
N
N N N N
Pd(PPh3)4 N
Cl Cl N N
N N

FIGURE 10.10  Ru(II) complex of tetradentate ligand. (Adapted from Zhang, G., R. Zong, H. W. Tseng, and
R. P. Thummel, Inorg. Chem, Angew., 47, 990–998, 2008. With permission.)

(H3cppd = 3,6-bis-(6′-carboxypyrid-2′-yl)-pyridazine) and [Ru2(cpph)(pic)4(µ-Cl)]+ (H2cpph = 1,4-bis(6′-


carboxypyrid-2′-yl)phthalazine) were also prepared by Xu et al. (2009).
The first family of mononuclear Ru–aqua WOCs trans-[Ru(pbn)(4-R-py)2(OH2)]2+ (pbn  =
2,2′-(4-(tert-butyl)pyridine-2,6-diyl)bis(1,8-naphthyridine); py = pyridine; R = Me, CF3, and NMe2) was
reported by Zong and Thummel (2005) for water oxidation. Zhang et al. (2008) (Figure 10.10) and
Tseng et al. (2008) have also reported a series of non-aqua Ru complexes that catalyze water oxida-
tion effectively. One class of those complexes, [Ru(dpp)(4-R-py)2]2+ (dpp = 2,9-dipyrid-2′-yl-1,10-
phenanthroline; R = Me, NMe2, and CF3), contain phenanthroline-based tetradentate ligands, which
readily bind ruthenium(II) in an equatorial tetradentate fashion and monodentate pyridyl ligands at
axial positions.

10.7.2 Iridium-Based Catalyst
Iridium oxide has long been recognized as a heterogeneous WOC with a low overpotential and
long-term stability (Youngblood et al. 2009). McDaniel et al. (2008) published the first family of Ir
WOCs, analogues of [Ir(ppy)2(OH2)2]+ (ppy = 2-phenylpyridine). Their TONs reported were impres-
sively high (e.g., 2760). However, catalytic rates were relatively low; thus, the reaction needed over
1 week to reach completion. Hull et al. (2009) changed one of the ppy ligands to Cp* (C5Me5) and
synthesized several [IrCp*(C^N)Cl] complexes (Figure 10.11).
As a step forward in developing Ir WOCs, they observed that the introduction of Cp* could
dramatically increase the catalytic rates of Ir-based WOCs. Unfortunately, these [IrCp*(C^N)Cl]
complexes showed lower TONs (1500 for [IrCp*(ppy)Cl]) than [Ir(ppy)2(OH2)2]+ and its analogues.
Then [IrCp*(C^N)(OH2)]+ was proposed as the real WOC with WNA on the Ir=O species responsi-
ble for the O–O bond formation (Blakemore et al. 2010). Further, Lalrempuia et al. (2010) modified
this complex and introduced an unusual pyridinium-carbene ligand instead of the ppy ligand. A new
complex was obtained, which displayed a remarkably high TON of ~10,000. They suggested that the

H2O N lr
Ce(IV) + Cl O2

FIGURE 10.11  Water splitting by Ir complex. (Adapted from Hull, J. F., D. Balcells, J. D. Blakemore, C. D.
Incarvito, O. Eisenstein, G. W. Brudvig, and R. H. Crabtree, J. Am. Chem. Soc., 131, 8730–8731, 2009. With
permission.)
Artificial Photosynthesis 203

high activity was due to the high electronic flexibility of the mesoionic pyridinium carbine ligand
which stabilizes the low oxidation state of the Ir complex with its neutral carbene-type resonance
form and the high oxidation state of the Ir complex with its charge-separated form.

10.7.3 Mn-, Co-, and Fe-Based Catalysts


Although nature has chosen Mn as the main building block of the catalytic OEC in the PS II, only
a handful of synthetic Mn complexes have been reported, which are capable of catalyzing water
oxidation. Thus, Mn(III,IV) dimer is capable of oxidizing water, when activated with a primary
oxidant. It is thought to be biomimetic in nature, because some of its structural and mechanistic
features are similar to those of the OEC. The catalytic mechanism of water oxidation based on the
Mn(III,IV) dimer was thought to involve the formation of a high-valent MnV=O or MnIV– O• species
(Mullins and Pecoraro 2008; Tagore et al. 2008). The oxyl radical is susceptible to nucleophilic
attack by a water molecule, forming a hydroperoxo intermediate that rapidly decomposes into O2
upon deprotonation. Besides Mn-based WOCs, some Co- and Fe-based WOCs were also discov-
ered. Interestingly, these catalysts contain only earth-abundant elements and are relatively stable
under catalytic conditions. Wasylenko et al. (2011) reported a Py5-Co complex [Co(Py5)(OH2)]2+
(Py5 = 2,6-(bis(bis-2′-pyridyl)methoxy-methane)-pyridine, which efficiently mediates the oxidation
of water electrochemically with a reaction rate of ~79 s−1. The nucleophilic attack on the Co(IV)-
hydroxy/oxo species by an incoming water/hydroxide substrate was suggested to be responsible for
the formation of the O–O bond (Figure 10.12).
Dogutan et al. (2011) discovered a mononuclear cobalt hangman corrole complex, which is capa-
ble of catalyzing water oxidation electrochemically. When immobilized in Nafion films, the TOF
for water oxidation at the single cobalt center of the hangman platform reached 0.81 s−1 driven at
1.4 V versus Ag/AgCl. The pendant –COOH group appears to benefit the O–O bond formation by
preorganizing the incoming water in close proximity to the cobalt oxo center. Nakamura and Frei
(2006) demonstrated that nanosized crystals of Co3O4 impregnated on mesoporous silica work effi-
ciently as oxygen-evolving catalysts. A wet-impregnation procedure was used to grow Co clusters
within the mesoporous Si as the template. The yield for clusters of cobalt oxide (Co3O4) nanosized
crystals was about 1600 times higher than for micron-sized particles, and the TOF was about 1140
oxygen molecules per second per cluster.
The first family of Fe(III)-tetraamido macrocyclic ligand WOCs (Fe(III)-TAMLs) was discov-
ered by Ellis et al. (2010). They reported that the oxygen evolution was rapid in the first 20 s and then
it became very slow. Only 16 turnovers were achieved for the best catalyst with TOFinitial = 1.3 s−1.
Both oxidative and hydrolytic decomposition pathways were proposed to limit the catalytic perfor-
mance of Fe(III)-TAMLs. Because iron is one of the most earth-abundant metals and is environ-
mentally friendly, this work will open up new avenues for the development of large-scale affordable
WOCs.

H
C O Aqua ligand to
Oxidatively stable
ligand environment N accommodate PCET

Co

Metal ion can accommodate Chelate effect conserves


multiple redox levels this Co-N bond

FIGURE 10.12  Catalytic water oxidation mediated by a high-valent cobalt complex. (Adapted from
Wasylenko, D. J., C. Ganesamoorthy, J. Borau-Garcia, and C. P. Berlinguette, Chem. Commun., 47, 4249–4251,
2011. With permission.)
204 Solar Energy Conversion and Storage

The realization of artificial systems that perform water splitting requires catalysts that produce
oxygen from water without the need for any excessive driving potential. Kanan and Nocera (2008)
reported an amorphous cobalt–phosphate catalyst that oxidizes water to O2. They reported that this
catalyst forms upon the oxidative polarization of an inert indium tin oxide electrode in phosphate-
buffered water containing cobalt(II) ions. The presence of phosphate was indicated in an approxi-
mate 1:2 ratio with cobalt in this material using various analytical techniques. The pH dependence
of the catalytic activity also implicates the hydrogen phosphate ion as the proton acceptor in the
oxygen-producing reaction. This catalyst not only forms in situ from earth-abundant materials but
also operates in neutral water under ambient conditions (Kanan et al. 2009). The use of earth-
abundant materials, operation in water at neutral pH, and formation of the catalyst in situ capture
functional elements of the OEC of photosystem II.

10.8  CATALYTIC PROTON REDUCTION AND HYDROGEN EVOLUTION

2H+ + 2e− ⎯Photosensitizer


hν,PRC
⎯⎯⎯⎯ → H2 (10.6)

The protons and electrons generated from catalytic water oxidation are combined to produce hydro-
gen. The process is assisted by proton reduction catalysts (PRCs).

10.8.1  Proton Reduction Catalyst by Hydrogenases


Hydrogenases are enzymes capable of catalyzing the oxidation of molecular hydrogen or its produc-
tion from protons and electrons according to the reversible reaction:

H2 ↽
!!! 2H + 2e
!⇀
+ − (10.7)

Most of these enzymes fall into two major classes: NiFe and Fe-only hydrogenases (Leonard et al.
1998). Extensive spectroscopic, electrochemical, and structural studies have been carried out to
explore the catalytic mechanism of hydrogenases. Although evolutionarily they are not related, the
NiFe and Fe-hydrogenases share a common, unusual feature, and that is an active-site low-spin Fe
center with CO and CN coordination.
Iron-iron hydrogenases (Nicolet et al. 2002; Peters 1999; Corr and Murphy 2011) are generally
found capable of reducing protons, although some of them are used for hydrogen oxidation and
bidirectional transformations. There are many different sources for molecular structures of Fe-Fe
hydrogenases. Many structural similarities were found among them, although their structures are
slightly different from each other under different crystallization states. Fe-Fe hydrogenases are well
known for their abilities to reduce protons to hydrogen, at nearly Nernstian potentials, where the
TOFs of Fe-Fe hydrogenase enzymes can reach a value of around 6000 mol of H2/mol per hydrog-
enase enzyme per second. Ni-Fe hydrogenases are primarily used for hydrogen uptake (Ogata et al.
2002; Casalot and Rousset 2001).
Hydrogenases are very old redox enzymes and frequently present in microorganisms belonging
to the Archaea and Bacteria domains of life; however, a few of them are found in Eukarya as well
(Brown and Doolittle 1997). They display remarkable performance on the reversible interconver-
sion between protons and hydrogen, since the purpose of hydrogenase enzymes is to facilitate a
charge separation or combination.
Ogata et al. (2002) reported the characterization of [NiFe] hydrogenase using Desulfovibrio
vulgaris. Characterization methods included x-ray crystallography and absorption and resonance
Raman spectroscopy. After the addition of CO, it was found to be bound to the Ni atom at the Ni-Fe
Artificial Photosynthesis 205

active site. The CO was not replaced with H2 in the dark at 100 K, but it was found to be liberated
by illumination with a strong white light. The Ni-C distances and Ni-C-O angles were about 1.77 Å
and 160°, respectively, except for one case (1.72 Å and 135°), where an additional electron density
peak between the CO and Sgamma (Cys546) was recognized. Distinct changes were observed in
the electron density distribution of the Ni and Sgamma (Cys546) atoms between the CO-bound and
CO-liberated structures for all the crystals tested. The novel structural features found near the Ni
and Sgamma (Cys546) atoms suggest that these two atoms at the Ni-Fe active site play a role during the
initial H2-binding process. Anaerobic addition of CO to dithionite-reduced [NiFe] hydrogenase led
to a new absorption band at about 470 nm (approximately 3000 M–1 cm–1). Resonance Raman spec-
tra (excitation at 476.5 nm) of the CO complex revealed CO-isotope-sensitive bands at 375/393 and
430 cm–1 (368 and 413 cm–1 for 13C18O). The frequencies and relative intensities of the CO-related
Raman bands indicated that the exogenous CO is bound to the Ni atom with a bent Ni-C-O structure
in solution.
Casalot and Rousset (2001) observed the high degree of similarity that exists between all
the [NiFe] hydrogenase operons. Near universality of hydrogen metabolism among microorgan-
isms suggests that the microbial ability to metabolize hydrogen is of great importance. The large
number of genes present in these operons are mostly involved in the maturation of the struc-
tural subunit, which is indicative of the complexity of the hydrogenase molecular structure. Two
main groups of maturation genes can be differentiated based on the resulting phenotypes, when
mutated:

• Cis-genes, encoding narrow specificity proteins. These are mainly located on the same
transcription unit as the structural genes.
• Trans-genes, encoding broad specificity proteins. These are located on different operons.
The maturation of the large subunit starts with the formation of a complex with the chap-
erone HypC, which remains bound to the amino terminus throughout processing. The
ligands CN and CO, which are derived from carbamoyl phosphate, are then inserted via
HypF and probably other accessory proteins. HypB is responsible for nickel atom delivery
in a GTP-hydrolysis-dependent reaction. The last identified step in the large subunit matu-
ration process is proteolytic cleavage at the carboxyl terminus.

The primary hydrogen binding site of Fe-hydrogenase has been identified by Nicolet et al. (2002).
An extensive genome sequencing effort has shown that eukaryotic organisms contain putatively
gene coding sequences that display significant homology to Fe-hydrogenases. A structural compari-
son was also carried out between Fe-hydrogenases and related proteins of unknown metal content
from yeast, plant, worm, insect, and mammals.
Peters et al. (1999) gave a clarification of the structures of iron-only hydrogenases from the
microorganisms Clostridium pasteurianum and Desulfovibrio desulfuricans and revealed that the
presumed site of reversible hydrogen oxidation exists as a unique, protein-associated organometallic
prosthetic group. The hydrogenase structures also provide insight into the chemical mechanism of
this highly evolved catalyst.
Corr and Murphy (2011) reported that hydrogenases catalyze redox reactions with molecular
hydrogen, either as substrate or product. The enzymes harness hydrogen as a reductant using metals
that are abundant and economical, namely, nickel and iron.
Interestingly, the [Fe]-hydrogenase is used by certain microorganisms in the pathway that reduces
carbon dioxide to methane. [Fe]-hydrogenase has consistently provided structural and mechanistic
surprises since its discovery two decades ago, often requiring complete reevaluation of its mecha-
nism of action.
Consequently, many synthetic catalysts have been inspired by the reactivity of hydrogenase
enzymes. Therefore, many variations of electrochemical and photochemical hydrogen production/
uptake systems based on Fe-complexes have been studied both experimentally and theoretically.
206 Solar Energy Conversion and Storage

Inspired by the catalytic action of hydrogenase enzymes, a variety of synthetic molecular cata-
lysts for hydrogen electro- and photochemical production have been developed. These catalysts are
mainly based on transition-metal centers incorporated to organic ligand frameworks. A substantial
number of studies have been conducted on cobalt (Krishnan and Sutin 1981; Hawecker et al. 1983;
Hu et al. 2005, 2007; Razavet et al. 2005; Du et al. 2008; Khnayzer et al. 2014; Sun et al. 2011),
nickel (Helm et al. 2011; Small et al. 2011), iron (Gloaguen et al. 2001; DuBois and DuBois 2008;
Kaur-Ghumaan et al. 2010), and molybdenum (Karunadasa et al. 2010; Merki and Hu 2011) com-
plexes, which have been shown to catalyze proton reduction in acidic acetonitrile or water to pro-
duce H2. Extensive investigations of the proton reduction using molecular catalysts provided deep
insights into the mechanism of this reaction. It was found that the proton reduction catalysis mainly
proceeds through the sequence of reduction-protonation steps, through the formation of metastable
metal–hydride intermediate, resulting in hydrogen evolution. Several interesting synthetic models
of hydrogenase have been developed by the group of DuBois (Helm et al. 2011). These Ni coordina-
tion complexes contained pendant base groups in the second coordination sphere of the metal cen-
ter, which resulted in tremendous enhancement of catalytic rates, which have reached an impressive
100,000 s–1 in wet acetonitrile. These complexes are considered as the closest synthetic models of
hydrogenase and one of the best electrocatalysts for the reduction of protons.

10.8.2 Light-Driven Catalysis for Hydrogen Production


Photochemical production of hydrogen from water can be considered to consist of two subsystems:

• Photochemical or electrochemical component, where the required oxidizing or reducing


equivalents are generated.
• Suitable redox catalysts assisting the formation of hydrogen and oxidation of water. Most
up-to-date efforts revolve around the second component, identifying suitable redox cata-
lysts. Noble metals such as Pt are known as good catalysts for H2 evolution for artifi-
cial systems but not for water oxidation. Electrolysis of water is best achieved using a Pt
electrode as cathode and a metal oxide such as RuO2 or IrO2 as the anode (Anthony and
Critchley 2005; Graetzel 1983).

Kiwi et al. (1982) discussed some basic features of light-induced electron transfer reactions in solu-
tions. Particular emphasis was placed on charge separation and kinetic control of the events by
organized molecular assemblies such as micelles and vesicles. They also defined some fundamental
aspects of redox catalysis with colloidal dispersions of noble metals and their oxides. The applica-
tion of these concepts to the problem of hydrogen and oxygen production from water by visible light
was illustrated. Methods to obtain highly active and selective catalysts were analyzed in detail, and
the performance of bifunctional redox catalysts in cyclic water cleavage systems by visible light has
been described.
Borgarello et al. (1981a) reported that zinc tetramethylpyridylporphyrin [(ZnTMPyP]4+ in acidic
aqueous solution sensitizes oxygen generation efficiently by visible light in the presence of acceptors
such as Fe3+ and Ag+ ions and colloidal RuO2/TiO2 redox catalyst. Hydrogen and oxygen are gener-
ated simultaneously under visible light illumination of [ZnTMPyP]4+ solutions, when a bifunctional
catalyst (Pt and RuO2 codeposited onto TiO2) was employed. Borgarello et al. (1981b) observed that
the conversion of light into chemical fuels in photochemical devices equipped with nonoxide semi-
conductor electrodes (e.g., n-CdS) is associated with a serious problem of photocorrosion, because
holes produced in the valence band of the semiconductor upon irradiation migrate to the surface of
the semiconductor, where photocorrosion occurs. It is remedied by a thin layer of RuO2 in microhet-
erogenous CdS systems. A CdS sol prepared in the presence of maleic anhydride/styrene copolymer
has been loaded with RuO2 and Pt. These CdS microelectrodes are surprisingly active catalysts for
cleavage of H2O and H2S.
Artificial Photosynthesis 207

The latter materials show very low overvoltage for water oxidation to molecular O2. A popular
and commonly used procedure in photochemical and electrochemical studies of H2 evolution is to
use a one-electron redox reagent like methyl viologen (4,4′-dimethyl-bipyridinium chloride, MV2+)
as an electron shuttle. MV2+ is very soluble in water and has a redox potential slightly more negative
than that of normal potential of hydrogen (E° = −0.44 V). This reagent is colorless in its oxidized
form but turns deep blue in singly reduced form (MV+). In the presence of suitable redox catalysts,
the reduced form is reoxidized readily with concomitant evolution of H2 from water as

MV2+ + e− → MV+ (10.8)

2MV+ + 2H+ → 2MV2+ + H 2 (10.9)

Photolysis of a [Ru(bpy)3]2+ complex with visible light in the presence of MV2+ (e− acceptor), a
sacrificial donor like EDTA disodium salt, and a redox catalyst leads to sustained evolution of H2
from water. Enzymes such as hydrogenases can also be used in biomimetic systems because the
photolysis can be carried out in neutral aqueous solutions.
Kohl et al. (2009) described a solution-phase reaction scheme that leads to the stoichiometric lib-
eration of dihydrogen and dioxygen in consecutive thermal-driven and light-driven steps mediated
by mononuclear ruthenium complexes. The initial reaction of water at 25°C with a dearomatized
ruthenium(II) pincer complex yields a monomeric aromatic Ru(II) hydrido–hydroxo complex,
which on further reaction with water at 100°C, releases H2 and forms a cis-dihydroxo complex.
Irradiation of this complex in the 320–420 nm range liberates oxygen and regenerates the start-
ing hydrido–hydroxo Ru(II) complex, most probably by elimination of hydrogen peroxide, which
rapidly disproportionates. Isotopic labeling experiments with H217O and H218O clearly show that
the process of an oxygen–oxygen bond formation is intramolecular, thus establishing a previously
elusive fundamental step toward dioxygen-generating homogeneous catalysis.
Many groups have shown that nanocrystals of catalytic metal oxides such as Ir-oxide or Nb-oxide
can be used efficiently as catalysts for the water splitting. Maeda et al. (2008) have demonstrated
that potassium hexaniobate nanoscrolls (NS-K4Nb6O17) formed by exfoliation of lamellar K4Nb6O17
can be used as redox catalysts for visible light–driven H2 production (λ > 420 nm) from water when
[Ru(bpy)3]2+ is photolyzed in the presence of EDTA as a sacrificial electron donor. After exposure
to sunlight, the photosensitizer [Ru(bpy)3]2+ reaches its excited state [Ru(bpy)3]2+* and is reduced to
[Ru(bpy)3]+ quickly by sacrificial electron donors like ascorbate. It returns back to [Ru(bpy)3]2+ after
transferring electrons to proton reduction catalysts, and the electrons transferred to the catalysts are
used to reduce protons to produce hydrogen.
Hoertz et al. (2007) reported that dicarboxylic acid ligands such as malonate, succinate, and
butylmalonate stabilize IrO2 particles (2 nm diameter) synthesized by hydrolyzing aqueous [IrCl6]2–
solutions. It was observed that monodentate (acetate) and tridentate (citrate) carboxylate ligands,
as well as phosphonate and diphosphonate ligands, are comparatively less effective as stabilizers
and lead to different degrees of nanoparticle aggregation, as confirmed by transmission electron
microscopy. Succinate stabilized 2 nm IrO2 particles are good catalysts for water photooxidation in
persulfate/sensitizer solutions. Ruthenium tris(2,2′-bipyridyl) sensitizers containing malonate and
succinate groups in the 4,4′-positions are also good stabilizers of 2  nm–diameter IrO2 colloids.
The excited-state emission of these surface-bound succinate-terminated sensitizer molecules is effi-
ciently quenched on a timescale of ~30 ns, most likely by electron transfer to Ir(IV). The excited
state of the bound sensitizer is quenched oxidatively on the timescale of ~9 ns in 1 M persulfate solu-
tions in pH 5.8 Na2SiF6/NaHCO3 buffer solutions. Electron transfer from Ir(IV) to Ru(III) occurs
with a first-order rate constant of 8 × 102 s–1 evolving oxygen. The TON for oxygen evolution under
these conditions was ~150. Thus, the sensitizer−IrO2 diad is a functional catalyst for photooxidation
of water, and it provides a useful building block for overall visible light water-splitting systems.
208 Solar Energy Conversion and Storage

10.9  CATALYSIS FOR CO2 REDUCTION AND FUEL PRODUCTION


Development of practical systems for converting CO2 through photocatalytic reduction processes to
useful chemicals using solar light is considered as one of the solutions to alleviate the twin problem
of global climate change and renewable energy utilization. Two approaches for photocatalytic CO2
reduction were reported: homogeneous reaction systems (mainly using transition-metal complexes)
and heterogeneous systems (mainly using inorganic semiconductor as light absorber).The reduction
of CO2 is one of the important reactions related to photosynthesis. There is an urgent need to find a
means of reducing CO2 gas to other C-1 products such as alcohols or methane. Some other reduc-
tion products such as formic acid (HCOOH) and carbon monoxide (CO) can also be utilized (Yui
et al. 2011).
While CO2-derived fuels such as methanol or methane can be used directly (e.g., in combustion
systems), some products require special use. For example, formic acid can be used inside a special
fuel cell to produce electricity. The side products of the formic acid fuel operation are carbon diox-
ide and water. A formic acid fuel cell may be a direct alternative to a hydrogen fuel cell. Formic
acid is advantageous as a renewable fuel because it is a liquid; therefore, it can be stored and treated
much easier than pressurized or liquid hydrogen. However, current methods of formic acid produc-
tion do not involve the use of renewable energy and require precious metals as catalysts as well as
harsh reaction conditions (i.e., high CO2/H2 pressure and high temperature). For example, the CO2
reduction to methane with hydrogen gas takes place in the presence of a transition-metal catalyst
and at high temperature and high pressure:

CO2 + 4H2 → CH4 + 2H2O  ΔG = −27 kcal/mol (10.10)

This reverse water gas shift (RWGS) reaction has become of commercial interest for the manu-
facturing of natural gas from the products of coal gasification. Its inverse reaction is called steam
reforming, which is useful for hydrogen production at the industrial level. Researchers have been
trying to find different methods of catalytic hydrogenation of carbon dioxide to methane at milder
conditions. There have been a number of efforts to reduce CO2 photocatalytically using Ru-bpy,
Re(CO)3bpy, metalloporphyrin complexes with or without additional metal/metal oxide catalysts.
This prompted extensive research on the development of new catalytic systems for photodriven
CO2 reduction.
Lehn and Ziesel (1982) reported that when the solutions of Ru(2,2′-bipyridine)32+, cobalt(II) chlo-
ride, and carbon dioxide in acetonitrile/water/triethylamine were irradiated by visible light, carbon
monoxide and hydrogen were generated simultaneously. This reaction involves photoinduced reduc-
tion of CO2 and H2O. They also reported that triethylamine served as electron donor in the Ru(2,2′-
bipyridine)32+/Co2+ system. The amount of gas produced and the selectivity ratio of CO/H2 depend
markedly on the composition of the system. When bipyridine was added, a decrease in CO genera-
tion and increase in H2 production were observed. With the addition of different tertiary amines,
NR3, both the quantity (CO + H2) and the ratio of CO/H2 increased along the sequence R = methyl,
ethyl, propyl. On using triethanolamine instead of triethylamine, higher selectivity for CO2 reduction
to CO occurred in preference to water reduction. Co2+ was found to be the most efficient mediator for
both CO and H2 generation and specifically promotes CO formation, whereas salts of other cations
yield only H2. The mechanism of the reaction may involve intermediate formation of Co(I) species.
These processes may be viewed as an abiotic photosynthetic system allowing simultaneous genera-
tion of CO and H2 and regulation of the CO/H2 ratio. The results obtained were also of significance
for solar energy conversion with consumption of a pollutant CO2.
Kalyanasundaram (1986) suggested that the excited state of tricarbonylchloro(polypyridyl)
rhenium(I) complexes, [Re(CO)3(LL)(Cl)] (LL=  2,2-bipyridine, 4,4′-dimethyl-2,2′-bipyridine,
1,10-phenanthroline, 5-chloro-1,10-phenanthroline, and 2,2′-bipyrazine), is emissive from the
metal-to-ligand charge-transfer (MLCT) state in solution at room temperature, and it undergoes
Artificial Photosynthesis 209

facile electron transfer reactions with a variety of electron donor and acceptor molecules. Three
aspects of the excited-state photophysics and photoredox chemistry were also presented on the
basis of laser photolysis studies: (1) sensitivity of the room temperature absorption and emission to
variations in the nature of the polypyridyl ligand and solvent; (2) excited-state absorption spectral
features; and (3) reversible and irreversible “reductive” quenching (using various amines as electron
donors) and their relevance to the photocatalytic reduction of CO2 to CO.
Kobayashi et al. (2014) observed the selective formation of dialkyl formamides through pho-
tochemical CO2 reduction. Photochemical CO2 reduction catalyzed by a [Ru(bpy)2(CO)2]2+/
[Ru(bpy)3]2+/Me2NH/Me2NH2+ system in CH3CN selectively produced dimethylformamide (DMF).
A ruthenium carbamoyl complex ([Ru(bpy)2(CO)-(CONMe2)]+) formed by the nucleophilic attack of
Me2NH on [Ru(bpy)2(CO)2]2+ in this process worked as the precursor to DMF. Thus, Me2NH acted
as both the sacrificial electron donor and the substrate, while Me2NH2+ functioned as the proton
source. Similar photochemical CO2 reductions using R2NH and R2NH2+ (R = Et, nPr, or nBu) also
afforded the corresponding dialkyl formamides (R2NCHO) together with HCOOH as a by-product.
It was reported that the main product from the CO2 reduction varies from R2NCHO to HCOOH with
AN increase in the alkyl chain length of the R2NH. The selectivity between R2NCHO and HCOOH
was found to depend on the rate of [Ru(bpy)2(CO)(CONR2)]+ formation.
The Re complexes were also explored in detail due to their advanced photophysical properties,
making these complexes valuable not only as catalysts, but also in applications such as sensors, opti-
cal switches, light-emitting devices, nonlinear optical materials, and radiopharmaceuticals (Kumar
et al. 2010; Coleman et al. 2008; Vlcek and Busby 2006). A major disadvantage associated with
these Re-based photocatalysts is that UV light is required for excitation due to their low absorbance
in the visible region. Therefore, solar-driven applications are not possible with these complexes until
their absorption maxima are shifted to visible range by some modifications.
Supramolecular systems for photocatalytic reduction of CO2 were developed to allow light absorp-
tion in the visible part of the spectrum and more efficient electron transfer. Heteromultinuclear
Ru–Re complexes for the photocatalytic reduction of CO2 were used by Gholamkhass et al. (2005).
Reithmeier et al. (2014) prepared mononuclear iridium (III) complexes [Ir(mppy)(tpy)X]
(mppy = 4-methyl-2-phenylpyridine, X = Cl, I) and binuclear analogues with various bis(2-phenyl-
pyridin-4-yl) bridging ligands. Kinetic measurements of the photocatalytic two-electron reduction
of CO2 to CO were investigated, and the influence of intermolecular interactions between two active
centers was observed. A comparison between the monometallic and the bimetallic complexes was
made, which indicated an enhanced lifetime (TON) of the covalently linked complexes, causing an
increased overall conversion of CO2.
It was observed that the proton-assisted, multielectron methods for CO2 reduction require much
less energy than the one-electron process to CO2/CO2− (E° = −1.9 V). It stimulated the use of these
multi-electron-transfer routes for CO2 reduction using transition-metal complexes. The most com-
mon electrochemical reactions leading to the reduction of CO2 together with the competing reaction
of hydrogen evolution follow. The electrochemical potentials of these redox processes are calculated
at pH 7 and presented versus NHE:

2H + + 2e− → H 2 E°ʹ = −0.41 V (10.11)

CO 2 +1e – → CO 2 − E°ʹ = –1.9 V (10.12)

CO 2 + 2H+ + 2e – → CO + H 2O E°ʹ = –0.52 V (10.13)

CO 2 + 2H+ + 2e – → HCOOH E°ʹ = –0.61 V (10.14)



CO 2 + 6H+ + 6e – → CH 3OH + H 2O E°ʹ = –0.38 V (10.15)

210 Solar Energy Conversion and Storage

CO 2 + 8H+ + 8e – → CH 4 + 2H 2O E°ʹ = –0.24 V (10.16)


Ishida et al. (1990) carried out an electrochemical reduction of CO2 catalyzed by [RuL1(L2)(CO)2]2+
[L1,L2 = (bipy)2, (bipy)(dmbipy), (dmbipy)2, or (phen)2], [Ru(phen)2(CO)Cl]+ (phen = 1,10-phenanth-
roline), and [RuL(CO)2Cl2] [L = 2,2′-bipyridine (bipy) or 4,4′-dimethyl-2,2′-bipyridine (dmbipy)]
by controlled potential electrolysis at –1.30 V versus saturated calomel electrode in acetonitrile–
water (4:1, v/v), MeOH, or MeCN–MeOH (4:1, v/v). They observed that there was no difference in
activities between various catalysts in acetonitrile–water (4:1, v/v). On introduction of the dmbipy
ligand in MeOH, the amount of produced carbon monoxide became larger than that of HCO2–. This
was attributed to the equilibrium constants among the reaction intermediates [RuL1(L2)(CO)2]2+,
[RuL1(L2)(CO)-{C(O)OH}]+, and [RuL1(L2)(CO)(CO2–)]+, which become smaller on substitution of
bipy by dmbipy, because of the donor property of the CH3 group.
Photochemical CO2 reduction is normally carried out in aqueous solutions or organic solvents
under 1 atm CO2 at room temperature. The concentration of dissolved CO2 is substantially higher
in organic solvents (0.28  M in acetonitrile) as compared to water (0.03  M). Many catalytic sys-
tems produce formate and CO as products; however, it was found that the formate-to-CO ratio
varied from system to system. Metallocarboxylates (also metallocarboxylic acids) were postulated
as intermediates in photochemical and electrochemical CO2 reduction and the water gas shift reac-
tion. Photolysis of particulate dispersions of TiO2 loaded with Ru catalyst in aqueous solutions in
the presence of CO2 leads to selective formation of methane at ambient temperatures (Yamashita
et al. 1998). However, reduction of water to H2 gas takes place in the absence of CO2. Photolysis of
aqueous dispersions of titania loaded with Cu catalyst has been found to yield methanol as a major
reduction product (Yui et al. 2011; Wu et al. 2005):

3
CO 2 + 2 H 2O ⎯Cu-TiO

⎯⎯⎯ → CH 3OH + O 2 (10.17)
2
2

Morris et al. (2009) have described CO2 reduction mediated by four classes of transition-metal
catalysts:

• Metal tetraaza-macrocyclic compounds


• Supramolecular complexes
• Metalloporphyrins and related metallomacrocycles
• Re(CO)3(bpy)X-based compounds

Carbon monoxide and formate were the primary CO2 reduction products. Bicarbonate/carbonate
production was also proposed.
Sutin et al. (1997) reported the transition-metal-based systems that generate hydrogen and/or
reduce carbon dioxide upon irradiation with visible light. Most of the systems involve polypyridine
complexes of the d6 centers cobalt(III), rhodium(III), iridium(III), ruthenium(II), and rhenium(I).
Complexes with diimine ligands serve as photosensitizers and/or catalyst precursors. The corre-
sponding d8 metal centers and d6 hydrides are important intermediates: bimolecular reactions of
the hydrides or their reactions with H2O/H3O+ are responsible for the formation of hydrogen. When
carbon dioxide is present, it may insert into the metal–hydride bond to yield formate. Mechanistic
schemes for some dual-acting photoconversion systems that generate both hydrogen and carbon
monoxide or formate were considered.
Kyle and Clifford (2014) presented the recent developments in the use of rhenium and manga-
nese bipyridine carbonyl catalysts for the electrochemical reduction of CO2. They described that
4,4′-tert-butyl-substituted complexes fac-Re(bpy-tBu)(CO)3X have been found to be more active
Artificial Photosynthesis 211

than the parent 2,2′-bipyridine complexes. It was observed that the presence of Bronsted acids
increases the activity of these catalysts, with stronger acids leading to more rapid catalysis.
Therrien et al. (2014) used a series of pyridine- and lutidine-linked bis-NHC palladium pin-
cer complexes for electrocatalytic CO2 reduction. Lutidine-linked complexes have shown catalytic
activity at potentials as low as −1.6 V versus Ag/AgNO3. According to DFT studies, the redox activ-
ity of the ligand contributed to the stability of the reduced species, potentially addressing a major
deactivation pathway of previous palladium pincer CO2 reduction electrocatalysts. Raebiger et al.
(2006) described a bimetallic palladium complex, {m-C6H4(triphos)2[Pd(CH3CN)]2}(BF4)4, for CO2
reduction. They observed that this catalyst exhibits high catalytic rates and larger turnover numbers.
La Porte et al. (2014) prepared a homogeneous, integrated chromophore/two-catalyst system,
which is capable of storing the light via photochemically driven reverse water gas shift reaction,
where the reducing equivalents are provided by renewable H2. The system consists of the chromo-
phore zinc tetraphenylporphyrin (ZnTPP), H2 oxidation catalysts of the form [Cp(R)Cr(CO)3]−, and
CO2 reduction catalysts of the type [Re(bpy-4,4′-R2)(CO)3]Cl.
Yuhas et al. (2011) presented a series of fully integrated porous materials containing Fe4S4 clus-
ters, dubbed “biomimetic chalcogels.” They examined the effect of third metal cations on the elec-
trochemical and electrocatalytic properties of the chalcogels and found that ternary biomimetic
chalcogels containing Ni or Co show increased activity in transformation of carbon dioxide. These
can be thought of as solid-state analogues of NiFe or NiFeS reaction centers in enzymes.
Chiericato et al. (2000) reported that transition-metal complexes of Fe, Co, and Ni incorporating
terdentate polyimine ligands derived from substituted DAPA (2,6-bis-[1-(phenylimino)ethyl]pyri-
dine) are effective electrocatalysts for the reduction of carbon dioxide. All the prepared complexes
exhibited some degree of electrocatalytic activity; however, this activity was strongly dependent
on the nature of the ligand, the metal center, and the nature of the redox process, whether it was
predominantly metal or ligand localized. [Ni(v-DAPA)2]2+ appeared to be the most active complex.
This material was able to reduce carbon dioxide electrocatalytically at ca.−1.0 V, which represents
a dramatic reduction of about 1 V in the overpotential.
Neri et al. (2015) prepared a low-cost nickel–cyclam complex covalently anchored to a metal
oxide surface. The role of the surface immobilization on enhancing the rate of photoelectron trans-
fer was confirmed using transient spectroscopy. [Ni(1,4,8,11-tetraazacyclo-tetradecane-6-carboxylic
acid)]2+ complex was shown to be an active electrocatalyst in solution for CO2 reduction.
A viable option for recycling carbon dioxide is through the sunlight-powered photocatalytic con-
version of CO2 and water vapor into hydrocarbon fuels over highly active nanocatalysts. Many
different approaches have been developed for the heterogeneous photocatalytic reduction of CO2
on TiO2, ZnO, and various other metal oxides. Rani et al. (2014) prepared (Cu, Pt)-sensitized TiO2
nanoparticle wafers for the photocatalytic conversion of CO2 and water vapor to hydrocarbon fuels.
Ehsan and He (2015) reported synthesis of common cation heterostructure via modification of zinc
oxide by zinc telluride (ZnTe) photocatalyst through a one-pot hydrothermal approach at a reaction
temperature of 180°C. The fabricated heterostructure consisted of ZnO flower-like nanostructures
(hundreds of nanometers for the rod length or sheet size of the petals and tens of nanometers for the
corresponding diameter or thickness) exhibited photocatalytic capability for the reduction of carbon
dioxide into methane under visible light irradiation (λ ≥ 420 nm). The difference in the activity of
materials with different morphology was explained by different exposed crystal planes of ZnO and
different surface area (15.0 and 5.6 m2 g–1 for sheet-like petals and rod-like petals, respectively).
Various modified metal oxides were also prepared for efficient photocatalytic reduction of CO2.
These metal oxides were modified by different methods, and several new efficient heterogeneous
catalysts (i.e., binary, ternary, and doped materials were used for reduction of CO2). Song et al.
(2015) synthesized ZnFe2O4/TiO2 heterostructured photocatalysts with different mass percentages
of ZnFe2O4 through the hydrothermal deposition method. The photocatalytic activities of the nano-
composites were tested by photocatalytic reduction of CO2 in cyclohexanol under UV light (main
wavelength at 360 nm) irradiation. It was shown that the main products were cyclohexanone and
212 Solar Energy Conversion and Storage

cyclohexyl formate. ZnFe2O4/TiO2 nanocomposites showed much higher photocatalytic perfor-


mance as compared with pure TiO2 and ZnFe2O4 samples.
Shown et al. (2014) prepared graphene oxide (GO) decorated with copper nanoparticles (Cu/GO)
and used these to enhance photocatalytic CO2 reduction under visible light. A rapid one-pot micro-
wave process was used to prepare the Cu/GO hybrids with various Cu contents. Metallic copper
nanoparticles (4–5 nm in size) in the GO hybrid significantly enhanced the photocatalytic activity
of GO, basically through the suppression of electron–hole pair recombination, reduction of bandgap
of GO and modification of its work function. X-ray photoemission spectroscopy studies indicated a
charge transfer from GO to Cu. Ehsan et al. (2014) demonstrated that ZnTe can be utilized as an effi-
cient catalyst for the photoreduction of CO2 into methane under visible light irradiation (≥420 nm).
The combination of ZnTe with SrTiO3 increased the formation of CH4 by efficiently promoting
electron transfer from the conduction band of ZnTe to that of SrTiO3 under visible light irradiation.
Yang and Jin (2014) synthesized Zn2GeO4 nanorods by a surfactant-assisted solution-phase route.
It was thought that the cetyltetramethylammonium cations (CTA+) preferentially adsorb on the
planes of Zn2GeO4 nanorods, leading to preferential growth along the c-axis to form the Zn2GeO4
rods with larger aspect ratio and higher surface area. Hence, photocatalytic activity for photoreduc-
tion of CO2 was improved. Jiang et al. (2014) prepared a series of novel microspheres of CdIn2S4
by hydrothermal process. CdIn2S4 was synthesized from l-cysteine, which exhibited higher photo-
catalytic activity for CO2 reduction and has a potential application for catalysis under visible light.
The mechanism of photocatalytic reduction of CO2 in methanol over CdIn2S4 was also proposed.
The narrow bandgap of the prepared catalyst promoted reduction of CO2 to dimethoxymethane and
methyl formate in methanol.
Li et al. (2014) synthesized a series of metal oxides (Ni/Zn/Cr layered double hydroxides [LDHs])
by coprecipitation and an annealing method at different temperatures. Their activities in photocata-
lytic reduction of CO2 with H2O vapor were tested at room temperature and atmospheric pressure.
It was reported that the sample calcined at 500°C possessed the highest catalytic activity with
respect to the formation of CH4 and CO as the major products. It was attributed to the interaction
of uniformly dispersed NiO, Cr2O3, ZnCr2O4, and NiCr2O4 with small grain size. Almeida et al.
(2014) prepared pure and Cr(III) and Mo(V)-doped BiNbO4 and BiTaO4 by the citrate method. Pure
BiNbO4 and BiTaO4 were obtained in triclinic phase at 600°C and 800°C, respectively. The metal
doping was found to strongly influence the crystal structure as well as the photocatalytic activity of
these oxides. They showed that Cr(III)-doped BiTaO4 and BiNbO4, in general, were more selective
for hydrogen production, while Mo(V)-doped materials were more selective for CO2 generation.
BiTaO4 showed higher photocatalytic activity than BiNbO4 for hydrogen production as well as for
CO2 generation. A negligible change of conduction band minimum potential (CBM) was found
for Mo(V)-doped materials, which indicates that there might be no improvement of the reduction
power of the material following the substitutional doping. However there is a slight shift of the CBM
potential, slightly increasing the reduction power in the case of Cr(III)-doped BiNbO4. This effect
was much stronger in the Cr(III)-doped BiTaO4.
Wang et al. (2014) synthesized mesoporous Fe-doped CeO2 catalysts with different Fe-doping
concentrations through a nanocasting route using ordered mesoporous SBA-15 as the template.
The samples were prepared by filling mesopores in silica template with a Fe-Ce complex precur-
sor, followed by calcination and silica removal. Then their catalytic activity was tested for the
photocatalytic reduction of CO2 with H2O under simulated solar irradiation. Fe species can effec-
tively enhance photocatalytic performance of the reduction of CO2 with H2O, when compared with
nondoped mesoporous CeO2 catalyst.
In nature, plants use photosynthesis to convert sunlight and carbon dioxide into oxygen and
carbohydrates. Plants use photosynthesis for producing food sources, and if one is able to mimic
this process, this can create a clean and affordable supply of renewable energy. However, if CO2
is to be used as the feedstock for the production of carbon-containing fuels, many scientific
and engineering challenges have to be overcome first. These challenges include an efficient and
Artificial Photosynthesis 213

inexpensive mechanism for capture of CO2 at the energy-generating sites or even the atmosphere.
The development of efficient, robust, and inexpensive catalysts for CO2 reduction is one of the
major scientific challenges. Finding ways to integrate CO2 reduction catalysis with water oxida-
tion half-reaction, all driven by the energy of sunlight presents another scientific and engineering
challenge.

REFERENCES
Abe, R., K. Hara, K. Sayama, K. Domen, and H. Arakawa. 2000. Steady hydrogen evolution from water
on eosin Y-fixed TiO2 photocatalyst using a silane-coupling reagent under visible light irradiation.
J. Photochem. Photobiol. A: Chem. 137: 63–69.
Abrahamsson, M. L. A., H. B. Baudin, A. Tran, C. Philouze, K. E. Berg, M. K. Raymond-Johansson, et al.
2002. Ruthenium-manganese complexes for artificial photosynthesis: Factors controlling intramolecu-
lar electron transfer and excited-state quenching reactions. Inorg. Chem. 41: 1534–1544.
Ahrens, M. J., L. E. Sinks, B. Rybtchinski, W. Liu, B. A. Jones, J. M. Giaimo, et al. 2004. Self-assembly
of supramolecular light-harvesting arrays from covalent multi-chromophore perylene-3,4:9,10-
bis(dicarboximide) building blocks. J. Am. Chem. Soc. 126: 8284–8294.
Albinsson, B., and J. Martensson. 2008. Long-range electron and excitation energy transfer in
donor-bridge-acceptor systems. J. Photochem. Photobiol. C. Rev. 9: 138–155.
Alibabaei, L., M. K. Brennaman, M. R. Norris, B. Kalanyan, W. Song, M. D. Losego, et al. 2013. Solar water
splitting in a molecular photoelectrochemical cell. Proc. Natl. Acad. Sci. USA. 110: 20008–20013.
Allakhverdiev, S. I. 2011. Recent progress in the studies of structure and function of photosystem II.
J. Photochem. Photobiol. B: Biol. 104: 1–8.
Allakhverdiev, S. I., V. D. Kreslavski, V. Thavasi, S. K. Zharmukhamedov, V. V. Klimov, T. Nagata, et al. 2009.
Hydrogen photoproduction by use of photosynthetic organisms and biomimetic systems. Photochem.
Photobiol. Sci. 8: 148–156.
Almeida, C. G., R. B. Araujo, R. G. Yoshimura, A. J. S. Mascarenhas, A. F. Da Silva, C. M. Araujo, et al.
2014. Photocatalytic hydrogen production with visible light over Mo and Cr-doped BiNb(Ta)O4. Int. J.
Hydrogen Energy. 39:1220–1227.
Alstrum-Acevedo, J. H., M. K. Brennaman, and T. J. Meyer. 2005. Chemical approaches to artificial photo-
synthesis. Inorg. Chem. 44: 6802–6827.
Anthony F. C., and C. Critchley (Eds.). 2005. Artificial Photosynthesis: From Basic Biology to Industrial
Applications, Wiley-VCH, Weinheim.
Aratani, N., D. Kim, and A. Osuka. 2009. Discrete cyclic porphyrin arrays as artificial light-harvesting
antenna. Acc. Chem. Res. 42: 1922–1934.
Bak, T., J. Novotny, M. Rekas, and C. C. Sorrell. 2002. Photo-electrochemical hydrogen generation from water
using solar energy. Material-related aspects. Int. J. Hydrogen Energy. 27: 991–1022.
Balzani, V., A. Credi, and M. Venturi. 1997. Photoprocesses. Curr. Opin. Chem. Biol. 1: 506–513.
Balzani, V., A. Credi, and M. Venturi. 2008. Photochemical conversion of solar energy. ChemSusChem. 1:
26–58.
Barber, J. 2009. Photosynthetic energy conversion: Natural and artificial. Chem. Soc. Rev. 38: 185–196.
Benniston, A. C. and A. Harriman. 2008. Artificial photosynthesis. Mater. Today. 11: 26–34.
Blakemore, J. D., N. D. Schley, D. Balcells, J. F. Hull, G. W. Olack, C. D. Incarvito, et al. 2010. Half-sandwich
iridium complexes for homogeneous water-oxidation catalysis. J. Am. Chem. Soc. 132: 16017–16029.
Borgarello, E., K. Kalyanasundaram, D. Duonghong, and M. Graetzel. 1981b. Cleavage of water by visible-
light irradiation of colloidal CdS solutions: Inhibition of photocorrosion by RuO2. Angew. Chem. Int.
Ed. 20: 987–988.
Borgarello, E., K. Kalyanasundaram, Y. Okuno, and M. Grätzel. 1981a. Visible light-induced oxygen genera-
tion and cyclic water cleavage sensitized by porphyrins. Helv Chim. Acta. 64: 1937–1942.
Brown, J. R., and W. F. Doolittle. 1997. Archaea and the prokaryote-to-eukaryote transition. Microbiol. Mol.
Biol. Rev. 61: 456–502.
Casalot, L., and M. Rousset. 2001. Maturation of the [NiFe] hydrogenases. Trends. Microbiol. 9: 228–237.
Chiericato, G. Jr., C. R. Arana, C. Casado, I. Cuadrado, and H. D. Abruña. 2000. Electrocatalytic reduction of
carbon dioxide mediated by transition metal complexes with terdentate ligands derived from diacetyl-
pyridine. Inorg. Chim. Acta. 300–302, 32–42.
Cogdell, R. J., T. H. P. Brotosudarmo, A. T Gardiner, P. M Sanchez, and L. Cronin. 2010. Artificial photosyn-
thesis—Solar fuels: Current status and future prospects. Biofuels. 1: 861–876.
214 Solar Energy Conversion and Storage

Coleman, A., C. Brennan, J. G. Vos, and M. T. Pryce. 2008. Photophysical properties and applications of Re(I)
and Re(I)–Ru(II) carbonyl polypyridyl complexes. Coord. Chem. Rev. 252: 2585–2595.
Collings, A. F., and C. Critchley (Eds.). 2005. Artificial Photosynthesis: From Basic biology to Industrial
Applications. Wiley-VCH, Weinheim.
Concepcion, J. J., J. W. Jurss, M. K. Brennaman, P. G. Hoertz, A. O. T. Patrocinio, N. Y. Murakami Iha, et al.
2009. Making oxygen with ruthenium complexes. Acc. Chem. Res. 42: 1954–1965.
Corr, M. J., and J. A. Murphy. 2011. Evolution in the understanding of [Fe]-hydrogenase. Chem. Soc. Rev. 40:
2279–2292.
Danielsson, R., M. Suorsa, V. Paakkarinen, P. A. Albertsson, S. Styring, E. M. Aro, et al. 2006. Dimeric
and monomeric organization of photosystem II. Distribution of five distinct complexes in the different
domains of the thylakoid membrane. J. Biol. Chem. 281: 14241–14249.
Dinner, B. A., and G. T. Babcock. 1996. Structure, dynamics and energy conversion efficiency in photosystem
II. In: Oxygenic Photosynthesis: The Light Reactions. D. R. Ort, C. F. Yocum, and I. F. Heichel (Eds.):
Kluwer: The Netherlands, pp. 213–247.
Dogutan, D. K., R. McGuire, and D. G. Nocera. 2011. Electrocatalytic water oxidation by cobalt(III) hangman
β-octafluoro corroles. J. Am. Chem. Soc. 133: 9178–9180.
Du, P., K. Knowles, and R. Eisenberg. 2008. A homogeneous system for the photogeneration of hydrogen from
water based on a platinum(II) terpyridyl acetylide chromophore and a molecular cobalt catalyst. J. Am.
Chem. Soc. 130: 12576–12577.
DuBois, M. R., and D. L. DuBois. 2008. The role of pendant bases in molecular catalysts for H2 oxidation and
production. C.R. Chim. 11: 805–817.
Ehsan, M. F., and T. He. 2015. In situ synthesis of ZnO/ZnTe common cation heterostructure and its visible-
light photocatalytic reduction of CO2 into CH4. Appl Catal B: Enviorn. 166–167: 345–352.
Ehsan, M. F., M. N. Ashiq, F. Bi, Y. Bi, S. Palanisamy, and T. He. 2014. Preparation and characterization of
SrTiO3-ZnTe nanocomposites for the visible-light photoconversion of carbon dioxide to methane. RSC
Adv. 4: 48411–48418.
Ellis, W. C., N. D. McDaniel, S. Bernhard, and T. J. Collins. 2010. Fast water oxidation using iron. J. Am.
Chem. Soc. 132: 10990–10991.
Fukuzumi, S. 2008. Development of bioinspired artificial photosynthetic systems. Phys. Chem. Chem. Phys.
10: 2283–2297.
Gersten, S. W., G. J. Samuels, and T. J. Meyer. 1982. Catalytic oxidation of water by an oxo-bridged ruthenium
dimer. J. Am. Chem. Soc. 104: 4029–4030.
Gholamkhass, B., H. Mametsuka, K. Koike, T. Tanabe, M. Furue, and O. Ishitani. 2005. Architecture of
supramolecular metal complexes for photocatalytic CO2 reduction: Ruthenium-rhenium bi- and tetra-
nuclear complexes. Inorg. Chem. 44: 2326–2336.
Gloaguen, F. D. R., J. D. Lawrence, and T. B. Rauchfuss. 2001. Biomimetic proton reduction catalyzed by an
iron carbonyl thiolate. J. Am. Chem. Soc. 123: 9476–9477.
Graetzel, M. (Ed.). 1983. Energy Resources through Photochemistry and Catalysis. Academic Press,
New York.
Gust, D., and T. A. Moore. 1989. Mimicking photosynthesis. Science. 244: 35–41.
Gust, D., D. Kramer, A. Moore, T. A. Moore, and W. Vermaas. 2008. Engineered and artificial photosynthesis:
Human ingenuity enters the game. MRS Bull. 33: 383–387.
Gust, D., T. A. Moore, and A. L. Moore. 2001. Mimicking photosynthetic solar energy transduction. Acc.
Chem. Res. 34: 40–48.
Hambourger, M., G. F. Moore, D. M. Kramer, D. Gust, A. L. Moore, and T. A. Moore. 2009. Biology and
technology for photochemical fuel production. Chem. Soc. Rev. 38: 25–35.
Hammarström, L., and S. Styring. 2008. Coupled electron transfers in artificial photosynthesis. Philos. Trans.
R. Soc. Lond. B Biol. Sci. 363: 1283–1291.
Harriman, A., and J. P. Sauvage. 1996. Strategy for constructing photosynthetic models: Porphyrin-containing
modules assembled around transition metals. Chem. Soc. Rev. 25: 41–48.
Hawecker, J., J. M. Lehn, and R. Ziessel. 1983. Efficient homogeneous photochemical hydrogen generation and
water reduction mediated by cobaloxime or macrocyclic cobalt complexes. Nouv. J. Chem. 7: 271–277.
Helm, M. L., M. P. Stewart, R. M. Bullock, M. R. DuBois, and D. L. DuBois. 2011. A synthetic nickel electro-
catalyst with a turnover frequency above 100,000 s–1 for H2 production. Science. 333: 863–866.
Herek, J. L., W. Wohlleben, R. J. Cogdell, D. Zeidler, and M. Motzkus. 2002. Quantum control of energy flow
in light harvesting. Nature. 417: 533–535.
Herrero, C., B. Lassalle-Kaiser, W. Leibl, A. W. Rutherford, and A. Aukauloo. 2008. Artificial systems related
to light driven electron transfer processes in PSII. Coord. Chem. Rev. 252: 456–468.
Artificial Photosynthesis 215

Hoertz, P. G., Y. Kim, W. J. Youngblood, and T. E. Mallouk. 2007. Bidentate dicarboxylate capping groups
and photosensitizers control the size of IrO2 nanoparticle catalysts for water oxidation. J. Phys. Chem.
B. 111: 6845–6856.
Hohmann-Marriott, M. F., and R. E. Blankenship. 2011. Evolution of photosynthesis. Ann. Rev. Plant Biol.
62: 515–548.
Hu, X., B. S. Brunschwig, and J. C. Peters. 2007. Electrocatalytic hydrogen evolution at low over-
potentials by cobalt macrocyclic glyoxime and tetraimine complexes. J. Am. Chem. Soc. 129:
8988–8998.
Hu, X., B. M. Cossairt, B. S. Brunschwig, N. S. Lewis, and J. C. Peters. 2005. Electrocatalytic hydrogen evolu-
tion by cobalt difluoroboryl-diglyoximate complexes. Chem. Commun. 37: 4723–4725.
Huang, P., A. Magnuson, R. Lomoth, M. Abrahamsson, M. Tamm, L. Sun, et al. 2002. Photo-induced oxida-
tion of a dinuclear Mn2II,II complex to the Mn2III,IV state by inter- and intramolecular electron transfer to
RuIII tris-bipyridine. J. Inorg. Biochem. 91: 159–172.
Hull, J. F., D. Balcells, J. D. Blakemore, C. D. Incarvito, O. Eisenstein, G. W. Brudvig, et al. 2009. Highly active
and robust Cp* iridium complexes for catalytic water oxidation. J. Am. Chem. Soc. 131: 8730–8731.
Huynh, M. H. V., D. M. Dattelbaum, and T. J. Meyer. 2005. Exited state electron and energy transfer in
molecular assemblies. Coord. Chem. Rev. 249: 457–483.
IEA (International Energy Agency). 2014. Key world energy statistics, IEA, pp. 1–82.
Ishida, H., K. Fukui, T. Ohta, K. Ohkubo, K. Tanaka, T. Terada, and T. Tanaka, 1990. Ligand effects of ruthe-
nium 2,2′-bipyridine and 1,10-phenanthrolinecomplexes on electrochemical reduction of CO2. J. Chem.
Soc. Dalton Trans. 2155–2160.
Jiang, W., X. Yin, F. Xin, Y. Bi, Y. Liu, and X. Li. 2014. Preparation of CdIn2S4 microspheres and application
for photocatalytic reduction of carbon dioxide. Appl Surf. Sci. 288: 138–142.
Kalyanasundaram, K. 1986. Luminescence and redox reactions of the metal-to-ligand charge-transfer excited
state of tricarbonylchloro-(polypyridyl)rhenium(I) complexes. J. Chem. Soc. Faraday Trans. II. 82:
2401–2415.
Kanan, M. W., and D. G. Nocera. 2008. In situ formation of an oxygen-evolving catalyst in neutral water con-
taining phosphate and Co2+. Science. 321: 1072–1075.
Kanan, M. W., Y. Surendranath, and D. G. Nocera. 2009. Cobalt–phosphate oxygen-evolving compound.
Chem. Soc. Rev. 38: 109–114.
Karunadasa, H. I., C. J. Chang, and J. R. Long. 2010. A molecular molybdenum-oxo catalyst for generating
hydrogen from water. Nature. 464: 1329–1333.
Kaur-Ghumaan, S., L. Schwartz, R. Lomoth, M. Stein, and S. Ott, 2010. Catalytic hydrogen evolution from
mononuclear iron(II) carbonyl complexes as minimal functional models of the [FeFe] hydrogenase
active site. Angew. Chem. Int. Ed. 49: 8033–8036.
Kelley, R. F., S. J. Lee, T. M. Wilson, Y. Nakamura, D. M. Tiede, A. Osuka, et al. 2008. Intramolecular energy
transfer within butadiyne-linked chlorophyll and porphyrin dimer-faced, self-assembled prisms. J. Am.
Chem. Soc. 130, 4277–4284.
Khnayzer, R. S., V. S. Thoi, M. Nippe, A. E. King, J. W. Jurss, K. A. El Roz, et al. 2014. Towards a comprehen-
sive understanding of visible-light photogeneration of hydrogen from water using cobalt(II) polypyridyl
catalysts. Energy Environ. Sci. 7: 1477–1488.
Kiwi, J., K. Kalyanasundaram, and M. Graetzel. 1982. Visible light induced cleavage of water into hydrogen
and oxygen in colloidal and microheterogeneous systems. Struct. Bond. 49: 37–125.
Kobayashi, K., T. Kikuchi, S. Kitagawa, and K. Tanaka. 2014. Selective generation of formamides through
photocatalytic CO2 reduction catalyzed by ruthenium carbonyl compounds. Angew. Chem. Int. Ed. 53:
11813–11817.
Kohl, S. W., L. L. Weiner, L. Schwartsburd, L. Konstantinovski, L. J. W. Shimon, Y. Ben-David, et al. 2009.
Consecutive thermal H2 and light-induced O2 evolution from water promoted by a metal complex.
Science. 324: 74–77.
Krishnan, C. V., and N. Sutin. 1981. Homogeneous catalysis of the photoreduction of water by visible light. 2.
Mediation by a tris(2,2′-bipyridine)ruthenium(II)-cobalt(II) bipyridine system. J. Am. Chem. Soc. 103:
2141–2142.
Kumar, A., S. S. Sun, and A. Lees. 2010. Photophysics and photochemistry of organometallic rhenium dii-
mine complexes. Top. Organomet. Chem. 29: 1–35.
Kyle, A. G., and P. K. Clifford. 2014. Recent studies of rhenium and manganese bipyridine carbonyl catalysts
for the electrochemical reduction of CO2. Chp-5. Adv. Inorg. Chem. 66: 163–185.
La Porte, N. T., D. B. Moravec, and M. D. Hopkins. 2014. Electron-transfer sensitization of H2 oxidation and
CO2 reduction catalysts using a single chromophore. Proc. Natl. Acad. Sci. 111: 9745–9750.
216 Solar Energy Conversion and Storage

Lalrempuia, R., N. D. McDaniel, H. Müller-Bunz, S. Bernhard, and M. Albrecht. 2010. Water oxidation cata-
lyzed by strong carbene-type donor-ligand complexes of Iridium. Angew. Chem. Int. Ed. 49: 9765–9768.
Lehn, J. M., and R. Ziessel. 1982. Photochemical generation of carbon monoxide and hydrogen by reduction
of carbon dioxide and water under visible light irradiation. Proc. Natl. Acad. Sci. U. S. A. 79: 701–704.
Leonard, C. J., L. Aravind, and E. V. Koonin. 1998. Novel families of putative protein kinases in bacteria and
archaea: Evolution of the “eukaryotic” protein kinase superfamily. Genome Res. 8: 1038–1047.
Li, B. J., Z. J. Wu, C. Chen, W. F. Shangguan, and J. Yuan. 2014. Preparation and photocatalytic CO2 reduction
activity of Ni/Zn/Cr composite metal oxides. J. Mol. Catal. 28: 268–274.
Lukas, A. S., Y. Zhao, S. E. Miller, and M. R. Wasielewski. 2002. Biomimetic electron transfer using low
energy excited states: A green perylene-based analogue of chlorophyll a. J. Phys. Chem. B. 106:
1299–1306.
Maeda, K., M. Eguchi, S. H. A. Lee, W. J. Youmgblood, H. Hata, and T. E. Mallouk. 2009. Photocatlytic hydro-
gen evolutionfrom hexaniobate nanoscrollsand calcium niobate nanosheetssensitized by ruthenium(II)
bipyridyl complexes. J. Phys. Chem. C.113: 7962–7969.
McConnell, I., G. Li, and G. W. Brudvig. 2010. Energy conversion in natural and artificial photosynthesis.
Chem. Biol. 17: 434–447.
McDaniel, N. D., F. J. Coughlin, L. L. Tinker, and S. Bernhard. 2008. Cyclometalated iridium(III) aquo
complexes: Efficient and tunable catalysts for the homogeneous oxidation of water. J. Am. Chem. Soc.
130: 210–217.
McEvoy, J. P., J. A. Gascon, V. S. Batista, and G. W. Brudvig. 2005. The mechanism of photosynthetic water
splitting. Photochem. Photobiol. Sci. 4: 940–949.
Melis, A. 2009. Solar energy conversion efficiencies in photosynthesis: Minimizing the chlorophyll antennae
to maximize efficiency. Plant Sci. 177: 272–280.
Merki, D., and X. Hu. 2011. Recent developments of molybdenum and tungsten sulfides as hydrogen evolution
catalysts. Energy Environ. Sci. 4: 3878–3888.
Meyer, T. J. 1990. Intramolecular control of excited state electron and energy electron transfer. Pure Appl.
Chem. 62:1003–1009.
Moore, T. A., D. Gust, P. Mathis, J. C. Mialocq, C. Chachaty, R. V. Bensasson, et al. 1984. Photodriven charge
separation in a carotenoporphyrinquinone triad. Nature. 307: 630–632.
Morris, A. J., G. J. Meyer, and E. Fujita. 2009. Molecular approaches to the photocatalytic reduction of carbon
dioxide for solar fuels. Acc. Chem. Res. 42: 1983–1994.
Mullins, C. S., and V. L. Pecoraro. 2008. Reflections on small molecule manganese models that seek to mimic
photosynthetic water oxidation chemistry. Coord. Chem. Rev. 252: 416–443.
Nakamura, R., and H. Frei. 2006. Visible light-driven water oxidation by Ir oxide clusters coupled to single Cr
centers in mesoporous silica. J. Am. Chem. Soc. 128: 10668–10669.
Naoki, A., D. Kim, and A. Osuka. 2009. Discrete cyclic porphyrin arrays as artificial light-harvesting antenna.
Acc. Chem. Res. 42: 1922–1934.
Neri, G., J. J. Walsh, C. Wilson, A. Reynal, J. Y. C. Lim, X. Li, et al. 2015. A functionalised nickel cyclam
catalyst for CO2 reduction: Electrocatalysis, semiconductor surface immobilisation and light-driven
electron transfer. Phys. Chem. Chem. Phys. 17: 1562–1566.
Nicolet, Y., C. Cavazza, and J. C. Fontecilla-Camps. 2002. Fe-only hydrogenases: Structure, function and
evolution. J. Inorg. Biochem. 91: 1–8.
Nocera, D. G. 2012. The artificial leaf. Acc. Chem. Res. 45: 767–776.
Ogata, H., Y. Mizoguchi, N. Mizuno, K. Miki, S. Adachi, N. Yasuoka, et al. 2002. Structural studies of the
carbon monoxide complex of [NiFe] hydrogenase from Desulfovibrio vulgaris Miyazaki F: Suggestion
for the initial activation site for dihydrogen. Am. Chem. Soc. 124: 11628–11635.
Ort, D. R., X. Zhu, and A. Melis. 2011. Optimizing antenna size to maximize photosynthetic efficiency. Plant
Physiol. 155: 79–85.
Peters, J. W. 1999. Structure and mechanism of iron-only hydrogenases. Curr. Opin. Struct. Biol. 9: 670–676.
Prashant, V. K., and D. Meisel. 2003. Nanoscience opportunities in environmental remediation. Compt. Rend.
Chimie. 6: 999–1007.
Raebiger, J. W., J. W. Turner, B. C. Noll, C. J. Curtis, A. Miedaner, B. Cox, and D. L. DuBois. 2006.
Electrochemical reduction of CO2 to CO catalyzed by a bimetallic palladium complex. Organometallics
25: 3345–3351.
Rani, S., N. Bao, and S. C. Roy. 2014. Solar spectrum photocatalytic conversion of CO2 and water vapour into
hydrocarbons using TiO2 nanoparticle membranes. Appl. Surf. Sci. 289: 203–208.
Razavet, M., V. Artero, and M. Fontecave. 2005. Proton electroreduction catalyzed by cobaloximes: Functional
models of hydrogenases. Inorg. Chem. 44: 4786–4795.
Artificial Photosynthesis 217

Reithmeier, R. O., S. Meister, B. Rieger, A. Siebel, M. Tschurl, U. Heiz, et al. 2014. Mono- and bimetallic
Ir(III) based catalysts for the homogeneous photocatalytic reduction of CO2 under visible light irradia-
tion. New insights into catalyst deactivation. Dalton Trans. 43: 13259–13269.
Röger, C., M. G. Müller, M. Lysetska, Y. Miloslavina, A. R. Holzwarth, and F. Würthner. 2006. Efficient
energy transfer from peripheral chromophores to the self-assembled zinc chlorin rod antenna: A bioin-
spired light-harvesting system to bridge the “green gap.” J. Am. Chem. Soc. 128: 6542–6543.
Romain, S., L. Vigara, and A. Llobet. 2009. Oxygen−oxygen bond formation pathways promoted by ruthe-
nium complexes. Acc. Chem. Res. 42: 1944–1953.
Ryan, D. M., M. K. Coggins, J. J. Concepcion, D. L. Ashford, Z. Fang, L. Alibabaei, et al. 2014. Synthesis and
electrocatalytic water oxidation by electrode-bound helical peptide chromophore-catalyst assemblies.
Inorg. Chem. 53: 8120–8128.
Sayama, K., K. Mukasa, R. Abe, Y. Abe, and H. Arakawa. 2002. A new photocatalytic water splitting sys-
tem under visible light irradiation mimicking a Z-scheme mechanism in photosynthesis. J. Photochem.
Photobiol. A: Chem. 148: 71–77.
Sens, C., I. Romero, M. Rodriguez, A. Llobet, T. Parella, and J. Benet-Buchholz. 2004. A new Ru complex
capable of catalytically oxidizing water to molecular dioxygen. J. Am. Chem. Soc. 126: 7798–7799.
Shown, I., H. C. Hsu, Y. C. Chang, C. H. Lin, P. K. Roy, A. Ganguly, et al. 2014. Highly efficient visible light
photocatalytic reduction of CO2 to hydrocarbon fuels by Cu-nanoparticle decorated graphene oxide.
Nano Lett. 14: 6097–6103.
Small, Y. A., D. L. DuBois, E. Fujita, and J. T. Muckerman. 2011. Proton management as a design principle
for hydrogenase-inspired catalysts. Energy Environ. Sci. 4: 3008–3020.
Song, G., F. Xin, and X. Yin. 2015. Photocatalytic reduction of carbon dioxide over ZnFe2O4/TiO2 nanobelts
heterostructure in cyclohexanol. J. Colloid Interfacial Sci. 442: 60–66.
Straight, S. D., G. Kodis, Y. Terazono, M. Hambourger, T. A. Moore, A. L. Moore, et al. 2008. Self-regulation
of photoinduced electron transfer by a molecular nonlinear transducer. Nat. Nanotechnol. 3: 280–283.
Sun, L., H. Berglund, R. Davydov, T. Norrby, L. Hammarstro, P. Korall, et al. 1997. Binuclear ruthenium–
manganese complexes as simple artificial models for photosystem II in green plants. J. Am. Chem. Soc.
119: 6996–7004.
Sun, L., L. Hammarström, B. Akermark, and S. Styring. 2001. Towards artificial photosynthesis: Ruthenium-
manganese chemistry for energy production. Chem. Soc. Rev. 30: 36–49.
Sun, Y., J. P. Bigi, N. A. Piro, M. L. Tang, J. R. Long, and C. J. Chang. 2011. Molecular cobalt pentapyridine
catalysts for generating hydrogen from water. J. Am. Chem. Soc. 133: 9212–9215.
Sutin, N., C. Creutz, and E. Fujita. 1997. Photo-induced generation of dihydrogen and reduction of carbon
dioxide using transition metal complexes. Comm. Inorg. Chem. 19: 67–92.
Tagore, R., R. H. Crabtree, and G. W. Brudvig. 2008. Oxygen evolution catalysis by a dimanganese complex
and its relation to photosynthetic water oxidation. Inorg. Chem. 47: 1815–1823.
Therrien, J. A., M. O. Wolf, and B. O. Patrick. 2014. Electrocatalytic reduction of CO2 with palladium bis-N-
hetercycliccarbene pincer complexes. Inorg. Chem. 53: 12962–12972.
Tseng, H. W., R. Zong, J. T. Muckerman, and R. Thummel. 2008. Mononuclear ruthenium(II) complexes that
catalyze water oxidation. Inorg. Chem. 47: 11763–11773.
Umena, Y., K. Kawakami, J. R. Shen, and N. Kamiya. 2011. Crystal structure of oxygen-evolving photosystem
II at a resolution of 1.9 Å. Nature. 473: 55–60.
Vlcek, A., and M. Busby. 2006. Ultrafast ligand-to-ligand electron and energy transfer in the complexes fac-
[ReI(L)(CO)3(bpy)]n+. Coord. Chem. Rev. 250: 1755–1762.
Wada, T., K. Tsuge, and K. Tanaka. 2000. Electrochemical oxidation of water to dioxygen catalyzed by the
oxidized form of the bis(ruthenium-hydroxo) complex in H2O. Angew. Chem. Int. Ed. 39: 1479–1482.
Wada, T., K. Tsuge, and K. Tanaka. 2001. Syntheses and redox properties of bis(hydroxoruthenium) com-
plexes with quinone and bipyridine ligands. Water-oxidation catalysis. Inorg. Chem. 40: 329–337.
Wang, Y., F. Wang, Y. Chen, D. Zhang, B. Li, S. Kang, et al. 2014. Enhanced photocatalytic performance of
ordered mesoporous Fe-doped CeO2 catalysts for the reduction of CO2 with H2O under simulated solar
irradiation. Appl. Catal. B: Environ. 147: 602–609.
Wasylenko, D. J., C. Ganesamoorthy, J. Borau-Garcia, and C. P. Berlinguette. 2011. Electrochemical evi-
dence for catalytic water oxidation mediated by a high-valent cobalt complex. Chem. Commun. 47:
4249–4251.
Wenger, O. S. 2009. Long-range electron transfer in artificial systems with d(6) and d(8) metal photosensitiz-
ers. Coord. Chem. Rev. 253: 1439–1457.
Wiberg, J., L. Guo, K. Pettersson, D. Nilsson, T. Ljungdahl, J. Mårtensson, et al. 2007. Charge recombination
versus charge separation in donor-bridge-acceptor systems. J. Am. Chem. Soc. 129: 155–163.
218 Solar Energy Conversion and Storage

Wu, J. C. S., H. M. Lin, and C. L. Lai. 2005. Photo-reduction of CO2 to methanol using optical-fiber photoreac-
tor. Appl. Catal. A: Gen. 296: 194–200.
Xu, Y., T. Åkermark, V. Gyollai, D. Zou, L. Eriksson, L. Duan, et al. 2009. A new dinuclear ruthenium com-
plex as an efficient water oxidation catalyst. Inorg. Chem. 48: 2717–2719.
Xu, Y., A. Fischer, L. Duan, L. Tong, E. Gabrielsson, B. Åkermark, et al. 2010. Chemical and light-driven
oxidation of water catalyzed by an efficient dinuclear ruthenium complex. Angew. Chem. Int. Ed. 49:
8934–8937.
Yamashita, H., Y. Fujii, Y. Ichihashi, S. G. Zhang, K. Ikeue, D. R. Park, et al. 1998. Selective formation of
CH3OH in the photocatalytic reduction of CO2 with H2O on titanium oxides highly dispersed within
zeolites and mesoporous molecular sieves. Catal. Today. 45: 221–227.
Yang, M., and X. Q. Jin. 2014. Facile synthesis of Zn2GeO4 nanorods toward improved photocatalytic reduc-
tion of CO2 into renewable hydrocarbon fuel. J. Cent. South Univ. 21: 2837–2842.
Yin, Q., J. M. Tan, C. Besson, Y. V. Geletii, D. G. Musaev, A. E. Kuznetsov, et al. 2010. A fast soluble carbon-
free molecular water oxidation catalyst based on abundant metals. Science. 328: 342–345.
Youngblood, W. J., S. H. A. Lee, K. Maeda, and T. E. Mallouk. 2009. Visible light water splitting using dye
sensitized oxide semiconductors. Acc. Chem. Res. 42: 1966–1973.
Yuhas, B. D., C. Prasittichai, J. T. Hupp, and M. G. Kanatzidis. 2011. Enhanced electrocatalytic reduction
of CO2 with ternary Ni-Fe4S4 and Co-Fe4S4-based biomimetic chalcogels. J. Am. Chem. Soc. 133:
15854–15857.
Yui, T., Y. Tamaki, K. Sekizawa, and O. Ishitani. 2011. Photocatalytic reduction of CO2: From molecules to
semiconductors. Topics Curr. Chem. 303: 151–184.
Zhang, G., R. Zong, H. W. Tseng, and R. P. Thummel. 2008. Ru(II) Complexes of tetradentate ligands related
to 2,9-di(pyrid-2′-yl)-1,10-phenanthroline. Inorg. Chem. 47: 990–998.
Zong, R., and R. P. Thummel. 2005. A new family of Ru complexes for water oxidation. J. Am. Chem. Soc.
127: 12802–12803.
11 Nanomaterials for Solar Energy
Mohammad Azad Malik, Sajid Nawaz Malik,
and Asma Alenad

CONTENTS
11.1 Introduction........................................................................................................................... 219
11.2 Binary Materials.................................................................................................................... 220
11.2.1 Copper Sulfide........................................................................................................... 220
11.2.2 Iron Sulfide................................................................................................................ 222
11.2.3 Lead Sulfide............................................................................................................... 223
11.2.4 Tin Sulfide................................................................................................................. 225
11.3 I–III–VI Materials................................................................................................................. 226
11.3.1 Copper Indium Disulfide (CuInS2)............................................................................ 227
11.3.2 Copper Indium Diselenide (CuInSe2)........................................................................ 229
11.3.3 Copper Indium Gallium Disulfide (CIGS)................................................................ 231
11.3.4 Copper Indium Gallium Diselenide (CIGSe)............................................................ 232
11.4 Copper-Zinc-Tin Chalcogenides (CZTSSe)........................................................................... 235
11.5 Copper Iron Tin Sulfide and Related Materials.................................................................... 242
References....................................................................................................................................... 245

11.1 INTRODUCTION
Global energy consumption during 2010 was recorded as 524 quadrillion British thermal units
(Btu). Various energy planning agencies have projected that this consumption will increase to 630
quadrillion Btu in 2020 and might rise to 820 quadrillion Btu in 2040 (U.S. Energy Information
Administration [USEIA] 2013). This growth in energy consumption is driven by increasing growth
in industrial activity in developing countries and changing lifestyles. Currently, most energy is
obtained by the burning of fossil fuels, mainly coal, oil, and natural gas. Other major sources of
energy are the nuclear reactors and hydroelectric power projects. The burning of fossil fuels results
in the emission of greenhouse gases, especially CO2 in the atmosphere, which has been identified
as a major cause of global warming. Many scientific studies indicate that a temperature increase of
1.8°C–4.0°C may occur in the global temperature during the twenty-first century, which may cause
serious and irreversible effects, such as the melting of polar ice, gigantic floods, and a rise in sea
level (Intergovernmental Panel on Climate Change [IPCC] 2007). The disposal of nuclear fuels
and nuclear security issues hinder the widespread use of nuclear energy. Large hydroelectric dams
suffer from the issues of seasonal flow fluctuations and periods of draught. Furthermore, obtain-
ing consent for more large hydroelectric sites is becoming increasingly difficult as such projects
are opposed by local communities and environmentalists due to associated ecological disruptions
(Sims 2008). Therefore, present energy supply trends are considered as being unsustainable for the
long term.
Now the focus is shifting toward alternative energy technologies that are more viable, clean,
efficient, and environmentally benign. The Sun is the most abundant source of energy. It produces
enormous amounts of energy due to thermonuclear fusion reactions. According to one estimate,
solar energy striking Earth in 1 h is more than the total energy consumed on the planet in a year

219
220 Solar Energy Conversion and Storage

(Dhere 2007). Another calculation suggests that the solar energy falling on Earth in 30 sunny days
is equivalent to the energy produced from all fossil fuels on Earth, either consumed or unused. If
efficient and economical means of harvesting this energy could be developed, all energy needs
could be satisfied by solar energy alone. Solar energy may, therefore, potentially ensure the transi-
tion toward a sustainable energy supply system for the twenty-first century.
Sunlight can be converted into useful energy by two routes: one is the solar thermal approach,
whereby solar energy is converted to heat; the second is the solar photovoltaic approach, which uses
semiconductor materials to generate electricity from solar radiation through the photoelectric effect.
The first viable solar electricity-producing cell was demonstrated by Chapin, Pearson, and Fuller of
Bell Telephone Laboratories in 1954 (Chapin et al. 1954). Since then this technology has shown a
steady and substantial growth predominantly driven by government subsidies and incentives. Many
policy papers on energy suggest that in 2030, solar photovoltaics may become a terawatt industry
satisfying a considerable fraction of total global energy consumption (Bellemare 2006). Despite all
this progress and inherent potentials of photovoltaics, the current share of photovoltaics in global
energy production is very low; the cost per watt–peak of electricity produced is still expensive. This
cost can be significantly reduced by market expansion, technological developments, improvement in
solar module efficiencies, and utilization of superior production strategies (Fthenakis 2009).
At the heart of the solar cell lies light-absorbing semiconductor material that absorbs photons
of light to produce electron and hole carriers via photovoltaic (PV) effect. The PV market is cur-
rently dominated by first-generation single-junction PV devices based on single or multicrystal-
line silicon wafers. However, cost per peak watt of electricity generated from a single-crystalline
silicon PV technology is not low enough to make it economically viable, and about half of the
cost involved is the material cost for silicon wafers. Second-generation PV technologies are thin-
film single-junction devices using semiconductors that offer the key advantage of reduced material
use and thus reduced associated costs, while maintaining the efficiencies comparable to or better
than first-generation PV technologies. This approach is based on thin layers of semiconductors
like cadmium telluride (CdTe), copper indium disulfide (CIS), copper indium gallium diselenide
(CIGSe), and copper-zinc-tin chalcogenides (CZTSSe) deposited onto low-cost substrates like glass.
These materials have high solar optical absorption coefficients (greater than 105 cm−1), so film thick-
ness is typically less than a micron and 100–1000 times lesser materials are required than silicon
wafers–based PV. Furthermore, these technologies have the potential for upscaling by roll-to-roll
manufacturing, thus offering even better economy of scales. Similarly, the third-generation photo-
voltaics encompass a range of evolving devices, upstarts, and wild ideas that have joined the race to
achieve the performance and cost goals desired for solar PV (Kazmerski 2006). These technologies
include polymer cells, biomimetics, quantum dot technologies, tandem/multijunction solar cells,
hot carrier cells, upconversion and downconversion technologies, and solar thermal technologies.
They may also include silicon nanostructures, modifying incident spectrum (concentration), use of
excess thermal generation (caused by UV light) to enhance voltages or carrier collection, and use of
infrared spectrum to produce electricity at night (Brown and Wu 2009).
The synthesis of nanocrystals, their properties, and their potential applications for solar energy
have been reviewed (Ramasamy et al. 2012, 2013a, 2013b; Akhavan et al. 2012; Abermann 2013;
Zhang et al. 2013a; Zhou et al. 2013; Aldakov et al. 2013; Fan et al. 2014; Azimi et al. 2014). This
chapter will cover various technological developments, especially the use of semiconductor nano-
materials in solar photovoltaics and their overall impact on the growth of this technology.

11.2  BINARY MATERIALS


11.2.1  Copper Sulfide
Copper sulfide exists in a number of stoichiometric and nonstoichiometric forms; therefore, it has
been an interesting material for fundamental studies. Its composition is based on nontoxic and
Nanomaterials for Solar Energy 221

Earth-abundant elements, and it finds important applications in solar photovoltaics, bioimaging, and
photocatalysis. Xiong and Zeng (2012) have reported the synthesis of multishelled copper sulfide
hollow spheres. Their approach involved the synthesis of polyvinylpyrrolidone (PVP)-coated Cu2O
spheres by polyol method and the carrying out of ion exchange of these Cu2O nanospheres using
thiourea and sodium sulfide, thus producing multishelled Cu2S hollow spheres. It was found that
the optical bandgap of the Cu2S hollow spheres varied upon increasing diameter of the spheres and
was found to be up to 2.10, 1.49, and to 1.42 eV for single, double, and triple-shelled Cu2S hollow
spheres, respectively.
Synthesis of hierarchical hollow spheres of CuS at the interface of water and oil has been
reported (Jiang et al. 2012a, 2012b). In a typical reaction, thioacetamide dissolved in water and cop-
per naphthenate dissolved in dimethylbenzene were allowed to react at the interface for 24 h at room
temperature. Amorphous hollow spheres were formed by the interfacial reaction at the interface of
water and dimethylbenzene. Hierarchical CuS hollow spheres were subsequently obtained by auto-
claving these amorphous spheres in ethanol at 60°C for 96 h.
Sun et al. (2012) have carried out the synthesis of polyhedral 26-facet Cu7S4 hollow cages by
using the sacrificial template method. Cu2O particles having 26 facets were synthesized as tem-
plates by reducing Cu(CH3COO)2 with glucose. The reaction of templates with Na2S at room tem-
perature resulted in the formation of Cu2O/Cu7S4 core–shell particles. Selective removal of Cu2O
core using ammonia yielded Cu7S4 hollow cages.
Synthesis of hybrid Ru-Cu2S nanostructures having cage and nanonet-like morphologies has
been reported by Vinokurov et al. (2012). Hybrid nanostructures are especially beneficial for solar
cell applications as they offer better electron–hole separation. Han et al. (2012) have developed a
one-pot colloidal method for the synthesis of Cu1.94S-ZnS, Cu1.94S-ZnS-Cu1.94S, and Cu1.94S-ZnS-
Cu1.94S-ZnS-Cu1.94S heteronanostructures. Copper iodide and zinc diethyldithiocarbamate complex
in oleyalamine were reacted to form heterostructures with screw-, dumbbell-, and sandwich-like
morphologies.
A wide variety of single-source precursors have been synthesized and utilized for preparation of
copper sulfide nanocrystals. Copper complexes of alkylxanthates, mercaptobenzothiazoles, thioben-
zoates, dithiocarbamates, and dithiolates have been synthesized and used for the synthesis of copper
sulfide nanocrystals with various sizes and morphologies. Abdelhady et al. (2012) have reported
the use of 1,1,5,5-tetra-iso-propyl-2-thiobiuret complex of copper as a single-source precursor for
the synthesis of copper sulfide nanocrystals in a continuous-flow microfluidic reactor. Spherical
Cu7S4 nanocrystals with an average diameter of 6.7 ±1.6 nm, 10.8 ± 1.9 nm and 11.4 ± 2.4 nm were
obtained at 170°C, 200°C, and 230°C, respectively. Thermolysis of a copper complex of S-methyl
dithiocarbamate in different high boiling solvents at different temperatures has been used for the syn-
thesis of copper sulfide nanocrystals (Bera and Seok 2012). High boiling solvents used in thermolysis
include ethylene glycol, ethylenediamine, hydrazine hydrate, and hexamethylenediamine. Sobhani
et al. (2012) have reported the use of [bis(thiosemicarbazide)copper(II)] chloride as a single-source
precursor for the preparation of copper sulfide nanoparticles through hydrothermal approach. CuS
nanoparticles synthesized by this method are 20–50 nm in size with an irregular morphology.
A low-temperature colloidal method has been used for the synthesis of ultrathin hexagonal cop-
per sulfides nanosheets (Du et al. 2012a). Crystallographic phase of the nanoparticles was deter-
mined by x-ray diffraction to be the covellite CuS phase. It was found that the nanoparticles have
a nanosheet-like morphology with a thickness of about 3.5 nm. CuS nanosheets obtained by this
method showed an absorption peak at 465 nm and emission peaks at 418 and 445 nm.
A controllable solvothermal synthesis of CuS with hierarchical structures has been reported
(Peng et al. 2014). Different morphologies like plates, nanoparticles, spheres, and nanoflowers have
been obtained by optimally varying the reaction parameters. Electrochemical characterization
results indicate that CuS structures exhibit remarkable capacitive performance. Flower-like CuS
demonstrated a high specific capacitance (597 Fg−1), excellent discharge rate, and a good stability,
which reflect significant potential of this electrode material for use in super capacitor devices.
222 Solar Energy Conversion and Storage

A simple wet chemical route for preparation of Cu2S nanoneedles by a room temperature reac-
tion of CuCl and Na2S has been developed by Kumarakuru et al. (2014). Formation of Cu2S nanon-
eedles takes place by the self-assembly of Cu2S nanoparticles, whereas thioglycerol (TG) has been
utilized as the capping agent. Formation of a polyphasic mixture containing different phases of
copper sulfide as well as copper oxide and copper chloride was indicated in powder x-ray diffraction
(p-XRD) studies. Deposition of copper sulfide thin films onto different substrates has been carried
out by employing the thermal evaporation technique (Saadeldin et al. 2014). X-ray diffraction stud-
ies demonstrate that an orthorhombic chalcocite (γ-Cu2S) phase has been deposited onto substrates.
Atomic force microscopy revealed that thin films composed of nanoparticles have an average size of
about 44 nm. Nanostructured assemblies of copper sulfide (CuS) have been synthesized by a facile,
template-free route (Kundu and Pradhan 2014). It was found that the CuS nanoplates or nanopar-
ticles underwent self-assembly to form either spheres or nanotubes. Detailed studies were carried
out to elaborate the mechanism for the formation of nanotubes.

11.2.2 Iron Sulfide
Iron sulfides have recently attracted significant interest as solar absorber materials, mainly due to
their suitable bandgap, high absorption coefficient, nontoxicity, and abundance in nature. Pyrite
nanocrystals have emerged as a promising absorber material for large-scale production of solar pho-
tovoltaic devices, chiefly because of their Earth-abundant and nontoxic nature. Significant research
interest has been devoted toward development of scalable routes for synthesis of phase-pure and
shape-controlled colloidal pyrite nanocrystals. However, surface defects in pyrite crystals are a
major detrimental factor for their use in solar cells. Iron sulfide nanocrystals having a variety of
sizes and morphologies have been synthesized by employing various synthetic methods. Synthesis
of FeS2 nanoplates by injecting an organometallic precursor [Fe(CO)5] into a solution containing
oleylamine and sulfur has been reported (Kirkeminde et al. 2012). Nanoplates thus prepared had a
lateral size of 150 nm and thickness around 30 nm. These irregular-shaped plates were composed
mainly of hexagonal crystallites. The absorbance spectra of FeS2 nanoplates grown for 180  min
exhibited an excitonic peak at 895 nm (1.38 eV), which corresponded to direct bandgap of FeS2. A
hybrid solar cell constructed by blending a 1:1 ratio of FeS2 nanoplates with P3HT demonstrated an
open-circuit voltage of 780 mV with a power conversion efficiency of 0.03%.
Steinhagen et al. (2012) have carried out the synthesis of iron sulfide (FeS2) nanocrystals having a
pyrite phase from solvent-based dispersions, or “solar paint,” for fabrication of photovoltaic devices.
Phase purity of the nanocrystals was demonstrated by the p-XRD and Raman spectroscopy. These
nanocrystals were spray-deposited onto substrates to form absorber layers in devices with different
architectures, including Schottky barrier, heterojunction, and organic/inorganic hybrid solar cells,
to evaluate their suitability as a photovoltaic material. None of the devices exhibited a PV response,
whereas the electrical conductivity of the nanocrystal films was about 4–5 S cm−1. This lack of PV
response may be attributed to the highly conductive surface-related defects in pyrite nanocrystals.
The synthesis of single-crystalline cubic iron pyrite (FeS2) nanowires has been carried out by
thermal sulfidation of steel foil (Cabán-Acevedo et al. 2012). Isolated nanocubes had a size of
~150 nm, whereas size of the dendrites was about 40 nm, and they were composed of smaller par-
ticles of ~10 nm. It was observed that the sizes of both the nanocubes and nanodendrites increased
by increasing the reaction time. The pyrite nanowires have length greater than 2 μm and a 4−10 nm
diameter. Electrical transport measurements showed the pyrite nanowires to be highly p-doped,
with an average resistivity of 0.18 ± 0.09 Ω cm and carrier concentrations of the order of 1021 cm−3.
Bandgap calculated by optical measurement was found to be around 0.9 eV.
Macpherson and Stoldt (2012) have reported the synthesis of pyrite nanocubes by reaction of
FeCl2 and elemental sulfur in different alkylamines. The nanocrystals obtained after initial heat-
ing at 250°C had a random oblate shape, whereas the first stage of growth at 200°C resulted in the
formation of nanocubes with large size distribution. If the reaction was further carried out to allow
Nanomaterials for Solar Energy 223

a second growth stage, the size distribution of the nanocubes was improved. The FeS2 nanocubes
prepared by this method had a lateral size of 37 ± 11 nm. Absorption measurements of the nanocu-
bes showed an indirect bandgap around 1.1 eV along with two excitonic transitions at 1.9 and 3.0 eV.
Morrish et al. (2012) have reported the preparation of FeS2 nanoparticles through plasma-assisted
sulfurization of Fe2O3 nanorods. Nanorods of Fe2O3 having an approximate size of 150 nm were
deposited by chemical bath deposition using FeCl3 and NaNO3 on fluorine-doped tin oxide (FTO)
glass plates. Formation of the marcasite phase was somewhat suppressed by prolonged sulfurization
of the F2O3 nanorods; however, complete eradication could not be achieved. The bandgap (direct) of
obtained FeS2 was found to be 1.2 eV.
Beal et al. (2011) have reported the synthesis of greigite (Fe3S4) nanoparticles with spherical
morphology and a size of 6.5  ±  0.5  nm by hot-injection approach. Magnetic properties of these
Fe3S4 nanoparticles were compared with magnetic properties of similar size Fe3O4 nanocrystals.
Greigite (Fe3S4) nanoparticles showed saturation magnetization of 12 emu g−1 at 10 K and blocking
temperature around ~50 K.
Greigite (Fe3S4) nanoparticles have been prepared by vapor–solid interaction using a laterally
resolving ultrahigh vacuum multimethod instrument (Bauer et al. 2014). Bandgaps of the FeS
nanoparticles synthesized using ethylene glycol, ethylenediamine, and ammonia were found to be
3.13 eV, 3.02 eV, and 2.75 eV, respectively (Maji et al. 2012). Photocatalytic activity of FeS nanopar-
ticles was determined by carrying out methylene blue degradation experiments; it was found to have
better activity than commercial TiO2.
Cummins et al. (2013) have reported the synthesis of phase-pure iron sulfide nanowires through
sulfurization of hematite nanowire arrays. Transmission electron microscope (TEM) images
showed hollow iron sulfide nanotubes with diameters in the range of 100−300 nm and wall thick-
nesses around 60 nm. The average length of the nanotubes was around 3 μm. Beal et al. (2012) have
recently reported the synthesis of Fe1–xS and Fe3S4 nanocrystals by reaction of elemental sulfur
with Fe(acac)2 in oleylamine at 200°C for 4 h. Thin sheets of Fe3S4 were formed by following these
conditions, whereas the same reaction carried out at above 300°C for 30 min resulted in formation
of Fe1–xS nanocrystals. The nanocrystals possessed a hexagonal plate and prism-like morphologies
with an average size of 70 nm. Hexagonal iron sulfide (Fe7S8) nanoflowers have been developed by
Wang et al. (2013) from thermal decomposition of ferric hexadecylxanthate at 260°C without any
solvent or inert gas protection, forming Fe7S8 nanoflowers.
Li et al. (2014) have carried out a systematic study of optical and electronic properties of pyrite
nanocrystal thin films. They had used a variety of ligands having different anchor and bridging
groups. It was found that the anchor group of ligand mainly controls the optical absorption. The
conductivity and photoconductivity of the nanocrystals are however, controlled by combined effects
of anchor and bridging groups. Zhu et al. (2014) have investigated the effect of reaction conditions
and the local chemical environment on the shape, composition, and crystallographic phase of iron
pyrite NCs synthesized by using the hot-injection approach. Different morphologies were obtained
by varying the solvents, concentration of reactants, and capping ligand (either trioctylphosphine
oxide or 1,2-hexanediol). These morphologies include short, branched, and chromosome-like rods
having ~10 nm diameter and 20–30 nm length as well as quasi-cubic NC agglomerates of 200 nm.
The as-synthesized iron pyrite NCs can be dispersed well in chloroform, chlorobenzene, toluene,
and hexane, and thus are promising in solution-processable photovoltaic applications.

11.2.3 Lead Sulfide
Lead sulfide is a narrow-bandgap semiconductor that has been extensively studied. Its direct band-
gap and large exciton Bohr radius make it a suitable material for various solar photovoltaic devices.
A variety of techniques have been employed for the synthesis of PbS nanocrystals. Khan et al.
(2013a) have used a hot-injection approach for the synthesis of nearly monodispersed PbS nanopar-
ticles with tunable bandgap. It was found that the size of the nanocrystals could be controlled by
224 Solar Energy Conversion and Storage

suitably varying the injection temperature and the reaction time. For example, nanoparticles grown
at 110, 120, 130, 150, and 160°C had a mean size of 2.3 ± 0.18, 2.7 ± 0.19, 3.5 ± 0.23, 5.5 ± 0.25, and
10 ± 0.3 nm, respectively. Bandgap values calculated from UV-Vis absorption spectra were found
to vary from ~1.7 eV for size of 2.3 ± 0.18 nm and ~0.6 eV for 10 ± 0.3 nm. Heterojunction solar
cells fabricated using these nanocrystals have demonstrated an open-circuit voltage of 235 mV for
10 ± 0.3 nm and 386 mV for 2.3 ± 0.18 size nanoparticles. The device constructed using nanopar-
ticles having bandgap of 1.2 eV exhibited the highest short-circuit current of 1.67 mA cm−2.
Synthesis of spherical PbS nanoparticles from PbCl2 and elemental sulfur using the hot-injection
approach has been reported (Nakashima et al. 2013). The experimental procedure involved the
injection of an oleylamine solution of sulfur into the reaction vessel containing PbCl2 and oleyl-
amine preheated at 70°C and 80°C, and the reaction was continued for different time periods rang-
ing from 1 to 25 min. The mean diameter of the particles grown at 70°C and 80°C was estimated to
be 4.47 ± 0.92 nm and 4.53 ± 0.57 nm for 1 min reaction, and 5.30 ± 0.52 nm and 6.27 ± 0.46 nm
for 15 min of reaction time. A red shift in the photoluminescence wavelength of the nanoparticles
from 1221 to 1288 nm was observed for the nanoparticles grown at 70°C from 1 to 15 min, whereas
the nanoparticles grown at 80°C demonstrated a red shift from 1282 to 1370 nm.
A wet chemical reaction using lead acetate and thiourea was used for the synthesis of self-
supporting arrays of quadrangular PbS nanopyramids (Hu et al. 2013a). The Scherrer equation was
used to calculate the crystallite size which was estimated to be about 150 nm. Scanning electron
microscope (SEM) images showed large sheets of 3–15 µm. These sheets were formed as a result of
self-assembly of 100–200 nm quadrangular nanopyramids.
Lead sulfide nanocrystals with a starfish like morphology were obtained by the reaction of lead
acetate and thioacetamide in the presence of cetyltrimethylammonium bromide (CTAB) and sodium
dodecyl sulfate (SDS) at 80°C (Li et al. 2013d). The selected area electron diffraction (SAED)
pattern from the star-shaped PbS crystals imaged along the [111] zone axis indicated that the arms
of the star-shaped PbS have grown along the [100] direction. The influence of various reaction
parameters like reaction temperature and lead source on the morphology of PbS nanocrystals has
been investigated.
Pan et al. (2013) have reported the synthesis of PbS quantum dots (QDs) in a flow reactor for
applications in solar cells. It was shown that the PbS quantum dots produced in the flow reactor had
a comparable performance to those synthesized in a batch reaction. A dual-temperature-stage flow
reactor was used to carry out the synthesis of PbS nanoparticles with optimal results. Precursor A
used in this method was composed of lead oxide, oleic acid (OA), and octadecene (ODE), whereas
precursor B contained bis(trimethylsilyl) sulfide (TMS) and ODE. The two precursors were mixed
together under nitrogen. The temperature of the reaction was controlled by thermocouples. The
reaction between the precursors at elevated temperature leads to the formation of PbS nuclei which
act as seed and subsequently grow to form PbS quantum dots. These quantum dots are finally iso-
lated by precipitation using acetone and are redispersed in toluene.
Multiple exciton generation (MEG) in PbS, and a PbSxSe1−x alloy have been studied (Midgett et al.
2013). It was observed that a linear decrease in MEG efficiency occurs for both PbS and PbSxSe1−x
alloyed dots with an increase in diameter within the strong confinement regime. Synthesis of PbS
nanoparticles has also been carried out as part of a composite material for use in solar cells and solar
water splitting (Kawawaki and Tatsuma 2013). For solar cell application, the PbS nanoparticles
are used in conjunction with plasmonic gold nanoparticles to serve as light-harvesting antennae,
whereas PbS nanoparticles have been decorated with Al-doped ZnO (AZO) nanorod arrays for solar
water-splitting applications (Hsu et al. 2013a).
A controlled synthesis of the high-quality flower-shaped PbS nanostructures by A solvothermal
approach using propylene glycol as solvent and a thio Schiff-base (2-(benzylideneamino)-benze-
nethiol) as a new sulfur source has been reported (Arani and Niasari 2014). It was demonstrated
that the good-quality, flower-shaped PbS nanoparticles having nearly uniform morphology and
size could be synthesized. Furthermore, the influence of various reaction parameters like solvents,
Nanomaterials for Solar Energy 225

reaction time, temperature, and concentration of sulfurizing agent on morphology and size of the
nanocrystals has been investigated.
Colloidal PbS quantum dots having very small size possess significant potential for increasing
the open-circuit voltages of quantum dot–based solar cells due to their large energy gap. A low-
temperature synthesis of ultra-small red light–emitting PbS QDs from organometallic precursors
in the presence of 1,2-dichloroethane has been developed by Reilly et al. (2014). Average size and
optical properties of the PbS quantum dots have been determined using HRTEM and optical spec-
troscopy. HRTEM images revealed that the PbS quantum dots have an average size of 1.6 nm with
a standard deviation of 0.2 nm.
PbS nanoparticle films have been deposited electrochemically onto ITO glass substrates (Mocanu
et al. 2014). Deposition experiments resulted in the formation of different morphologies of PbS
nanostructures in the presence of different water-soluble polymers like polyacrylic acid (PAA),
PVP, polyvinyl alcohol (PVA), and poly(2-acrylamido-2-methylpropane sulfonic acid) (PAMPSA).
SEM images revealed that at least two different morphologies of PbS nanostructures are deposited
for all samples. The role of polymers in determining the shape of PbS nanostructures has been
explained.
McPhail and Weiss (2014) have investigated the reaction mechanism for the chemical reaction
of elemental sulfur with 1-octadecene (ODE). They have described the factors that induce a change
in morphology of the PbS quantum dots from cubic to hexapodal during synthetic reaction between
the S/ODE precursor and lead(II) oleate. Hexapodal geometry is obtained when the organosulfur
ligands bind to the growing QDs causing a preferential growth at the ⟨100⟩ faces rather than at ⟨111⟩
faces. It was also suggested that S/ODE can be used more reliably by decreasing the reaction tem-
perature and time for dissolution of the sulfur.

11.2.4  Tin Sulfide


Tin sulfide belongs to the family of IV–VI semiconductors. Three forms of tin sulfide exist, which
include SnS, SnS2, and Sn2S3. The bandgaps of SnS, SnS2, and Sn2S3 are 1.3, 2.18, and 0.95  eV,
respectively. Significant research attention has been devoted to explore the use of tin sulfide in solar
photovoltaic devices as an absorber based on Earth-abundant elements and having a low toxic-
ity. SnS has been reported to be either a p-type or n-type conductor depending on the tin content.
Furthermore, its conductivity might also change upon heat treatment. SnS2 is an n-type semiconduc-
tor, whereas Sn2S3 possesses highly anisotropic conduction behavior.
The synthesis of nanocrystalline tin sulfide is mostly carried out with a view to synthesize the
SnS phase, having a quantum confinement effect. The synthesis of SnS nanocrystals having three
different morphologies, such as cubes, spherical polyhedral, and nanosheets, has been reported by
Biacchi et al. (2013). Spherical polyhedral nanoparticles synthesized by this method had a mean size
of 9.7 ± 1.5 nm. SnS nanoparticles with cubic morphology were obtained by following the same pro-
cedure but carrying out injection of SnS at 170°C and maintaining the reaction temperature between
165°C and 170°C for 1 h. Resulting nanocubes had an average edge length of 11.5  ± 1.9 nm. A
slightly modified procedure for the preparation of SnS nanosheets has also been reported by the
same group. They have used a thermolytic reaction of tin acetate and elemental sulfur dissolved in
oleylamine. Thus-prepared nanosheets had a lateral dimension of about 270 ± 50 nm. Investigations
into the photocatalytic activity of these SnS nanocrystals against methylene blue dye showed that
the SnS with nanocube morphology had the highest photocatalytic activity.
Diethylthiocarbamate complex of tin has been used as a single-source precursor for the
phase-controlled synthesis of nanosized SnS, SnS2, and SnS/SnS2 heterostructures (Hu et al. 2013b).
Thermolysis of the complex in oleylamine at 320°C for 30 min yielded SnS nanocrystals, whereas
nanocrystals with SnS2 phase were obtained by using 300 µL CS2 in addition to the precursor during
thermolysis. Liang et al. (2013) have reported the synthesis of SnS nanocrystals by thermal decom-
position of SnO and elemental sulfur in oleic acid and oleylamine at different reaction temperatures.
226 Solar Energy Conversion and Storage

The size of the nanocrystals increased from 50 to 200 nm when the decomposition temperature was
increased from 150°C to 210°C. Similarly, morphology of the nanocrystals was changed from spherical
to sheet-like morphology by increasing the volume of oleic acid used in the reaction. A hydrothermal
reaction of SnCl2.2H2O, hydrazine, and thiourea at 180°C for 23 h has been used for the preparation of
SnS nanorods (Iqbal et al. 2013). The lengths of thus-prepared nanorods varied from 1 to 2 µm, whereas
the width was estimated to be about 80 nm. It was demonstrated that as-prepared SnS nanorods had a
maximum hydrogen absorption value of 0.73 wt%, demonstrating their potential as a hydrogen storage
material.
Yan et al. (2013a) have carried out microwave-assisted hydrothermal synthesis of CuS, ZnS, and
SnS nanocrystals and compared it with conventional hydrothermal synthesis. SnS nanoparticles
prepared by both processes, conventional hydrothermal heating and microwave hydrothermal heat-
ing, had nearly the same morphology; however, microwave-assisted heating proved to be much
quicker (15 min heating period) than conventional heating (4 h).
Sonochemical preparation of SnS nanoparticles exhibiting quantum confinement has been
reported (Azizian-Kalandaragh et al. 2013). In a typical reaction, an aqueous solution of tin chloride,
sodium sulfide, and triethanolamine was sonicated for 2 h at room temperature by a high-intensity
ultrasonic transducer. Polydispersed nanoclusters having particle sizes smaller than 100 nm were
thus obtained. The bandgap value of the SnS nanoparticles was determined to be 1.74 eV, indicating
the presence of the quantum confinement effect.
Rath et al. (2014) have proposed a fabrication scheme for improving the device performance of
low-cost chalcogenide-based solar cells. They have developed an In2S3 layer having embedded SnS
quantum dots. Their fabrication scheme involves incorporation of this intermediate bandgap layer
in copper indium sulfide (CIS) cells for improving the current generation and efficiency. They have
also proposed measures to optimally cap the surface of quantum dots for effective passivation of
defects and to protect the quantum dots from doping from the environment.

11.3  I–III–VI MATERIALS


Polycrystalline copper indium/gallium sulfur/selenium (CuIn1–xGa xS1–ySey) thin films have attracted
considerable research and industrial interest among chalcogenide semiconductor photoabsorber
materials because of their high optical absorption coefficient (105  cm–1), tunable bandgap energy,
long-term photoirradiation stability, and higher efficiency as compared to other absorber materi-
als. CIGS thin-film solar cells have already exceeded 20% power conversion efficiency and are
approaching the performance of silicon solar cells. A power conversion efficiency of 20.4% has
been reported for CIGS-based thin-film solar cells, whereas solar devices based on potassium-
doped CIGS have demonstrated an efficiency of 20.8%. Typical configuration of a CIGS-based solar
device is glass substrate/Mo back contact/CIGS-based absorber layer/CdS or ZnS buffer layer/TCO
window layer/antireflection coating layer (MgF2).
The most critical process that governs the cost and performance of CIGS solar cells is the deposi-
tion of the photoabsorber layer. So far, the CIGS thin films with a suitable composition profile, phase
purity, and grain structure have generally been deposited by the use of vacuum-based techniques
like co-evaporation or sputtering followed by post-treatment operations like high-temperature
annealing. However, these methods cannot be scaled up for mass production and require huge capi-
tal investment on vacuum and deposition equipment. Nonvacuum methods, in contrast, offer the
possibility of faster, scaled-up, continuous, roll-to-roll manufacturing and a low initial capital cost.
The scarcity of materials, such as In and Ga, has resulted in increased costs and has raised con-
cerns about the availability of these metals for large-scale deployment of CIGS-based solar cells.
Therefore, efficient processes making more judicious use of such materials are required to reduce
the material costs and to overcome the barriers to commercialization. These methods rely on solu-
tion-based deposition of CIGS nanoparticles or molecular precursors at low temperature followed
by post-treatment at higher temperature.
Nanomaterials for Solar Energy 227

CuInS2, CuInSe2, CuInGaS2 (CIGS), and CuInGaSe2 (CIGSe) are the leading I–III–VI2 mate-
rials finding application in solar photovoltaics. During the past two decades, extensive research
studies have been undertaken to develop controllable synthesis of I–III–VI2 materials as nanocrys-
tals. These nanocrystals are then used for low-cost, solution-based nonvacuum deposition of the
photoabsorber layer for thin-film solar cells. Subsequent sections will deal with the synthesis of
nanoparticles of these materials, which can be used as inks for fabrication of a photoabsorber layer
in solar cells.

11.3.1  Copper Indium Disulfide (CuInS2)


Copper indium disulfide (CuInS2) is an important ternary chalcogenide material of the I–III–VI
class of compound semiconductors, owing to its direct bandgap value of 1.5  eV closely corre-
sponding with the solar spectrum, excellent photoirradiation stability, and high absorption coef-
ficient (>105). CuInS2 was regarded as the material of choice for thin-film solar photovoltaics
(Yan et al. 2013b). Solar cells based on its quaternary selenium analogue CuIn xGa1–xSe2 have
already demonstrated submodule power conversion efficiency of 20.4% (Jackson et al. 2011).
Theoretical calculations suggest power conversion efficiencies of 27%–32% for CuInS2-based
solar cells (Connor et al. 2009). However, recombination losses in the space charge region have
limited the practical efficiencies to only 13% (Zhong et al. 2008). Other important applications
of CuInS2 nanostructures include in bioimaging (Liu et al. 2013c), for H2 evolution from water
(Tsuji et al. 2006), in light-emitting diodes (Chen et al. 2013b), and as counterelectrode material
for dye-sensitized solar cells (Yao et al. 2013).
Nanoparticles of AgInS2 and CuInS2 have been synthesized by hot injection at 270°C using cop-
per acetate, indium acetate, and silver acetate as the metal sources (Nadar et al. 2013). Dodecanethiol
and elemental sulfur were used as sulfide precursors, whereas oleylamine was used as surfactant.
It was demonstrated that the formation of the desired material is slow in dodecanethiol, probably
due to its dual role as surfactant and sulfur source. Binary phases could be traced in the reaction
products even after 4 h of reaction. In contrast, the samples synthesized using elemental sulfur
had no/little traces of intermediate binary phases such as β-In2S3. Generalized phase-controlled
synthesis of zinc blende and wurtzite CuInS2, Cu2SnS3, and Cu2ZnSnS4 nanoparticles has been car-
ried out using a wet chemical reaction (Chang and Waclawik 2013). Synthesis of CuInS2 involved
the use of copper iodide, indium acetate, and dodecanethiol in either octadecene or oleylamine. It
was observed that the nature of the coordinating solvent and temperature control the evolution of
the crystallographic phase of the nanoparticles. Triangular pyramidal nanocrystals of zinc blende
phase with an average size of 9.3 ± 0.5 nm were formed when noncoordinating octadecene was used
as solvent, whereas wurtzite CuInS2 nanoparticles were obtained when octadecene was replaced
by oleylamine. The optical bandgap of zinc blende and wurtzite CuInS2 was found to be 1.39 and
1.50 eV, respectively. A similar effect of the ligand on the phase of nanoparticles was also observed
in preparation of Cu2SnS3 and Cu2ZnSnS4 nanoparticles.
Chen et al. (2013a) have demonstrated the in situ growth of CuInS2 nanocrystals onto nano-
porous TiO2 film by solvothermal treatment. They have thoroughly investigated the effect of the
precursor’s concentration on the morphology of as-grown nanostructures. The CuInS2 film formed
on the nonconductive glass side was removed by scraping, whereas the film on the nanoporous
TiO2 side was repeatedly washed with deionized water and absolute ethanol. Highly ordered potato
chip–like arrays were formed with the 0.01 M and 0.03 M concentration of InCl3 used in the reac-
tion, whereas the formation of flower-shaped structures with an average diameter 3 μm was found to
cover the entire FTO/compact-TiO2/nanoporous-TiO2 substrate when the concentration of InCl3 was
increased to 0.1 M. The pores of nanoporous-TiO2 film were also filled by CuInS2 nanoparticles.
This later film was used to fabricate a heterojunction solar cell of configuration FTO/TiO2/CIS/
P3HT/PEDOT:PSS/Au which yielded a power conversion efficiency of 1.4%. Further optimization
of the CuInS2 layer as well as cell structure might improve the efficiency of the cell.
228 Solar Energy Conversion and Storage

Guo et al. (2013) have reported the preparation of inorganic ligand–capped CuInS2 nanocrystals
by a ligand exchange reaction using organic ligand (dodecanethiol and oleylamine)–capped CuInS2
nanoparticles and (NH4)2S. These sulfide (S2–)-capped nanoparticles were subsequently used as
aqueous ink to fabricate counterelectrodes by drop casting the ink onto cleaned FTO glass sub-
strates. Thin films were obtained by drying at room temperature and sintering at 500°C for 30
min in an argon environment. These films were used as counterelectrode in dye-sensitized solar
cells (DSSCs). The power conversion efficiency of the DSSCs significantly improved from 0.35%
to 6.32% with the use of inorganic ligand–capped CuInS2 nanocrystals as counterelectrode in the
DSSC. A further improvement to 6.49% was observed when the nanocrystals were sintered at
500°C. This efficiency value is comparable to that of platinum (Pt) counterelectrodes and demon-
strates a possibility of substituting costly Pt electrodes with low-cost CuInS2 nanocrystals–based
counterelectrodes.
Metastable wurtzite and zinc blende forms of CuInS2 nanoparticles were synthesized by the
one-pot reaction of copper-thiourea precursors having chloride, sulfate, and nitrate counterions
and indium sulfate/indium acetate in ethylene glycol (Gusain et al. 2013). Optical bandgap of the
wurtzite CuInS2 nanoparticles was found to be 1.4 eV. Studies were carried out to determine the
effect of incorporation of Ga3+ and Fe3+ ions into a CuInS2 lattice. It was observed that the wurtzite
structure was retained by replacing small amounts of In3+ by Ga3+ atoms. Higher loading of Ga3+
atoms into CuInS2 results in the formation of mixed chalcopyrite and wurtzite phases. The band-
gap value calculated from the UV-visible diffuse reflectance spectrum for the iron-doped samples
exhibited a decrease from 1.40 to 1.05 eV.
The spray pyrolysis technique has been used to deposit polycrystalline thin films composed
of CuInS2 nanocrystals with the size range of 40–60 nm onto glass substrates with a bandgap
of 1.55 eV. A solar device fabricated by using these films demonstrated a power conversion effi-
ciency of 7.60% (Khan et al. 2013b). Wurtzite CuInS2 nanowires have been synthesized by means
of Ag2S nanocrystals catalyzed growth in a solution-phase reaction (Li et al. 2013a). Typical syn-
thesis was carried out by using diethyldithiocarbamate complexes of silver, copper, and indium.
Optical and photoelectrical measurements of thin films prepared by drop casting as-synthesized
nanowires showed promising photoresponse characteristics. Bandgap of the nanowires was cal-
culated to be 1.5 eV. A similar solution-based approach has also been used for the synthesis of
single-crystalline wurtzite ternary CuInS2 and quaternary semiconductor CuIn xGa1–xS2 nanorib-
bons (Li et al. 2013b). It was observed that Cu1.75S nanocrystals formed in the initial reaction
stage serve as a catalyst for anisotropic growth of the nanoribbons. The optical bandgaps of the
as-synthesized CuIn xGa1–xS2 nanoribbons could be varied from 1.44 to 1.91 eV by varying the Ga
concentration.
Monodispersed CuInS2 nanopompons and hierarchical nanostructure have been synthesized by
a solvothermal route using Cu2O and In(OH)3 as metal precursors, thioacetic acid as sulfur source,
and ammonia (Liu et al. 2013d). These films were subsequently used as counterelectrodes of dye-
sensitized solar cells. A power conversion efficiency of up to 4.8% demonstrated good catalytic
activity of as-synthesized nanoflake films as counterelectrodes in DSSCs. Yang et al. (2013) have
also used the same method using different concentrations of CuSO4.5H2O and InCl3.H2O metal
precursors and CH3CSNH2 as the sulfur source in ethanol as the solvent. SEM imaging revealed
that the as-deposited films are composed of vertically aligned nanosheets. TEM images show that
the nanosheets consist of a large number of crystal grains. Yao et al. (2013) have synthesized CuInS2
nanocrystals capped by organic ligand dodecanethiol to prevent aggregations and to control the size
and morphology. A complete phase transfer was observed by mixing and stirring the two solutions,
thus yielding S2−-capped CuInS2 nanoparticles in the polar formamide phase. Thin films of both
the organic-capped as well inorganic S2−-capped CuInS2 nanocrystals were prepared. These films
were then used as counterelectrodes in DSSCs. It was observed that significant improvement occurs
in the optoelectronic properties of CuInS2 nanoparticles by replacement of organic surfactant with
all inorganic S2− ligand.
Nanomaterials for Solar Energy 229

Highly luminescent CuInS2 nanoparticles were synthesized by a noninjection method (Yu et


al. 2013). Reaction conditions were optimized by using different copper and indium compounds
as metal sources, varying molar ratios of Cu, In, and SDPP, and employing different solvents/cap-
ping ligands like octadecene, oleic acid, and oleylamine. The CuInS2 nanocrystals showed high
crystallinity with a mean diameter of 3.4 ± 0.4 nm.
The CuInS2 particles were synthesized by using Cu(NO3)2 and In(NO3)3 as the metal precursors
and CS(NH2)2 as the sulfur source in an aqueous solution at 160°C for 2 h (Yidong et al. 2013).
The particles thus prepared were deposited as thin films onto the quartz substrate by a spin-coating
process. Optical bandgap energy of the films was determined to be 1.40 eV. A heating-up synthe-
sis approach using Cu(acac)2, In(acac)3, elemental sulfur, and octadecene was used (Dierick et al.
2014). TEM images showed that the as-synthesized chalcopyrite nanocrystals had a quasi-spherical
morphology with a mean diameter of ~8 nm. No peaks from the binary phases were detected in the
p-XRD pattern, and the bandgap of the nanocrystals was found to be 1.54 eV.
Deng et al. (2014) have recently reported a generalized strategy for the synthesis of a variety
of ternary metal sulfides with controlled size, morphology, crystallographic phase, and stoichio-
metric composition. Various materials synthesized by this approach include orthorhombic Cu3BiS3
nanosheets and nanoparticles, orthorhombic Cu4Bi4S9 nanowires and nanoribbons, wurtzite CuInS2
nanopencils, cubic AgBiS2 nanocubes, orthorhombic Ag8SnS6 nanoparticles, and orthorhombic
Cu3SnS4 nanorods.

11.3.2  Copper Indium Diselenide (CuInSe2)


CuInSe2 nanowires were synthesized by a solution–liquid–solid (SLS) growth process catalyzed
by gold–bismuth core–shell nanoparticles (Wooten et al. 2009). CuInSe2 nanowires have also been
grown by thermal decomposition of single-source precursor [(PPh3)2Cu(μ-SePh)2In(SePh)2] and by
using multiple source precursors. It was observed that the morphology and stoichiometric composi-
tion of the nanowires obtained were strongly influenced by the chemical nature of the precursors Use
of single-source precursor [(PPh3)2Cu(μ-SePh)2In(SePh)2] leads to the formation of single-crystalline,
straight, and stoichiometric CuInSe2 nanowires, whereas optimization of the reaction parameters was
required to grow good-quality CuInSe2 nanowires by using multiple source precursor systems.
Thin films of CuIn(Se,S)2 have been fabricated by using a solution-processed, hydrazine-based
approach (Liu et al. 2009). CuInSe2 films were deposited onto various substrates including Mo-coated
glass and thermally oxidized silicon wafers by spin coating without carrying out annealing in a
toxic Se- or S-containing environment. CuInSe2 films with good crystallinity could be obtained
by simple heat treatment in an inert environment. It was possible to tailor bandgap of the absorber
layer by varying the amount of sulfur in the film. Solar cells having configuration glass/Mo/CIS/
CdS/i-ZnO/ITO were constructed using thus-prepared hydrazine-processed CuInSe2 absorber layer,
which demonstrated a power conversion efficiency of 12.2% under AM 1.5 illumination. Xu et al.
(2010) have developed a facile chemical approach for preparation of morphology-controlled CuSe,
CuInSe2 nanowire, and CuInSe2/CuInS2 core–shell nanocable bundles. Cubic Cu2–xSe nanowire
bundles were first synthesized and used as a self-sacrificial template for preparation of hexagonal
CuSe nanowire with a diameter of about 10–15 nm and lengths up to hundreds of micrometers. It
was observed that the smaller copper ions diffuse outward from the interior to the surface of nanow-
ires and subsequently react with sulfur and indium ions to form a CuInS2 shell over CuInSe2 core.
These nanocable bundles exhibited strong optical absorption in UV–vis region.
Ahn et al. (2012) have fabricated a CuInSe2 particles–based absorber layer for solar cells using
binary nanoparticles as precursors through a nonvacuum route. CuInSe2 absorber thin films were
obtained by selenization of the nanoparticle-coated film for 30 min. It was observed that nonuni-
form growth of the particles occurred, leaving large voids in the final films when these films are
subjected to selenization. These voids are detrimental to the performance of the solar cell as they
acted as short-circuiting paths in the solar cells. An additional solution-filling treatment was applied
230 Solar Energy Conversion and Storage

to the binary precursor film to mitigate this issue, and a conversion efficiency of up to 1.98% from
a CuInSe2 film selenized at 430°C was obtained.
Malik et al. (2011) have synthesized CuInSe2, CuGaSe2, and CuIn1–xGa xSe2 (CIGS) nanoparticles
by thermolysis of the diisopropyldiselenophosphinatometal complexes Mx[iPr2PSe2]n (M = Cu(I),
In(III), Ga(III); n = 1, 3) in HDA/TOP at 120°C –210°C or 250°C. The diameters of the nanoparti-
cles for CuInSe2, CuGaSe2, and CuIn0.7Ga0.3Se2 were found to be 4.9 ± 0.6 nm (at 180°C), 13.5 ± 2.9
nm (at 250°C), and 14 ± 2.22 nm (at 250°C), respectively. They had previously reported the prepara-
tion of CuInSe2 nanoparticles by a two-step reaction using CuCl, InCl3, and TOPSe in TOPO as the
coordinating solvent and ligand (Malik et al. 1999).
A simple heating-up of reactants approach has been described for high-yield synthesis of the
chalcopyrite phase of quaternary CuIn(S1–xSex)2 nanoparticles (Chiang et al. 2011). Gradual heating
of the reactants to a temperature of 265°C and maintaining this temperature led to the formation
of CuIn(S1–xSex)2 nanoparticles with controllable S/Se ratios. Tuning of the bandgap energies of the
nanocrystals from 0.98 to 1.46 eV was also feasible.
Stolle et al. (2012) have compared the photovoltaic response of the absorber layer based on
CuInSe2 nanocrystals capped with organic ligand (oleylamine) and inorganic ligands like metal
chalcogenide–hydrazinium complexes (MCCs), S2−, HS−, and OH−. Thin-film solar cells were fab-
ricated with CuInSe2 nanocrystals capped with different capping ligands used as absorber layer. It
was observed that the PV device based on MCC ligand-capped CuInSe2 nanocrystal demonstrated
a power conversion efficiency of 1.7% under AM 1.5 illumination, while the PV device based on the
oleylamine-capped CuInSe2 nanocrystals exhibited 1.6% under similar conditions. Hsin et al. (2011)
have demonstrated the feasibility for transformation of phase change material In2Se3 to CuInSe2
through a solid-state reaction. This study therefore represented a unique practical process for trans-
formation of the phase change material In2Se3 to the solar energy material CuInSe2 and the prepa-
ration of the nanoheterostructures composed of In2Se3/CuInSe2 for use in future nanodevices and
solar cells.
A method based on microwave irradiation assisted chemical reaction for preparation of multi-
phase CuInSe2 nanoparticles from copper acetate, indium acetate, and elemental selenium has been
reported (Ahn et al. 2010). This method also featured formation of a metastable CuSe phase that
enabled fabrication of a solution-processed, crack-free, crystalline, and high-performance CuInSe2
absorber layer for a solar cell. Solar cells with configuration glass/Mo/CISe/CdS/i-ZnO/n-ZnO/Al
were then fabricated by using either CuSe phase–free or CuSe phase–containing CuInSe2 nanopar-
ticles to assess the role of the CuSe phase on the preparation of a device-quality absorber layer. It
was demonstrated that a CuInSe2 absorber layer fabricated from CuSe containing CuInSe2 mul-
tiphase nanoparticles showed a power conversion efficiency of 8.2%, while open-circuit voltage,
short-circuit current density, and fill factor were measured as 0.44  V, 33.7  mA cm−2, and 55%,
respectively.
A simple method for colloidal preparation of highly luminescent CuInSe2 nanocrystals through a
silylamide-promoted reaction has been reported (Yarema et al. 2013). CuInSe2 nanocrystals of aver-
age sizes ranging from 2.7 and 7.9 nm and small size distribution were prepared by suitably varying
the reaction temperature and time.
A unique approach for activating the elemental selenium for use in the synthesis of metal sel-
enide nanocrystals via the solvothermal method has been described (Zhang et al. 2014). Higher
reactivity of Se2− ions than zero-valent Se atoms also contributes toward improvement of reaction
kinetics. Oleksak et al. (2014) have used a one-pot microwave-assisted solvothermal reaction for the
synthesis of CuInSe2 nanoparticles. The reaction was typically carried out by using a combination
of precursors, which strongly absorb microwave energy, and low microwave–absorbing solvents
like tri-n-octylphosphine (TOP) and oleic acid.
Lim et al. (2013) have deposited CuInSe2 thin films from nanoparticle precursors synthesized by
a solution-based colloidal approach. Absorber films with ca. 20% more density were obtained by
cold-isostatic pressing (CIP). It was demonstrated that thus formed CuInSe2 thin films had improved
Nanomaterials for Solar Energy 231

microstructure, lower porosity, a more uniform surface morphology, and a relatively thinner MoSe2
layer. A significant increase in photovoltaic performance of the solar cells fabricated by using these
films was observed. A threefold increase in the average efficiency, from 3.0% to 8.2%, was also
demonstrated.
CuInSe2 nanoparticles have been synthesized by using a two-step, continuous-flow, solar micro-
reactor (Kreider et al. 2014). Radiative heat from simulated, concentrated solar radiation was
used as a faster heating source, which served to reduce the reaction duration. Synthesis of both
chalcopyrite and sphalerite phases of CuInSe2 nanoparticles was carried out by suitably altering
the nucleation temperature and residence time in the solar microreactor. It was observed that the
formation of the chalcopyrite phase of CuInSe2 nanoparticles is favored by the higher nucleation
temperatures and longer residence times. A continuous method has also been reported for synthe-
sis of CuInSe2 nanoparticles in a microtubular reactor (Kim et al. 2014). Monodispersed colloidal
CuInSe2 nanoparticles thus formed were used as nanocrystal ink for deposition of CuInSe2 thin
films. The shape of the nanocrystals gradually transformed from spherical to hexagonal to trigonal
with increasing In or Se content. CuInSe2 nanoparticles synthesized at a high temperature pos-
sessed trigonal morphologies with chalcopyrite crystallographic structures. Utilization of these inks
for solar photovoltaics was verified by fabricating a lab-scale device with 1.9% efficiency under
AM 1.5 G illumination.
A hybrid ink has been prepared from copper-rich CuSe nanoparticles and an indium precursor
solution to form CuInSe2 thin films for solar cell applications (Cho et al. 2014). PV devices fabri-
cated by using these films demonstrated a power conversion efficiency of 5.04% as compared to
an efficiency of 1.04% for normally synthesized copper-rich CuSe nanoparticles. This observation
confirmed that the Cu-MEA complex had a strong influence on the performance of CuInSe2-based
solar cells produced with the hybrid ink process.

11.3.3  Copper Indium Gallium Disulfide (CIGS)


Pan et al. (2009) have reported the synthesis of quaternary Cu1.0Ga xIn2–xS3.5 and
Cu1.0In xTl2–xS3.5 nanocrystals by hot-injection approach. TEM images showed nearly monodis-
persed Cu1.0Ga xIn2–xS3.5 nanocrystals with an average diameter of 6.2  nm. Bandgap of alloyed
Cu1.0Ga xIn2–xS3.5 nanocrystals, as determined by UV-vis absorption spectroscopy, could be tailored
in the range of 1.43 to 2.42 eV depending upon the composition.
CuInGaS2 (CIGS) nano-ink has been prepared using the hot-injection approach (Singh et al.
2012). CIGS nanocrystals were obtained by precipitation with ethanol and centrifugation at 10,000
rpm for 20 min. Nano-ink prepared by dispersing CIGS nanocrystals in toluene is used to fabricate
thin films by drop casting onto glass substrates. Optical bandgaps of CuInS2 thin films were found
to be 1.57 eV, which was slightly higher than bulk CuInS2 bandgap because of the small size of
nanocrystals (~5 nm). CIGS thin films exhibited a bandgap of 1.65 eV as a result of Ga substitution.
An et al. (2014) have reported a simple strategy based on paste coating and sulfurization for
fabrication of CIGS thin films for photovoltaic applications. The films prepared using this paste
demonstrated good photoelectrical characteristics; however, multiple coating and drying cycles
were required to deposit thin films of the desired thickness. Another paste was prepared by using
ethanol as solvent and ethyl cellulose as organic binder. Solar devices with configuration Al/Ni/
ZnO:Al/i-ZnO/CdS/CIGS/Mo-coated glass were fabricated to compare the performance of both
pastes. The solar cell fabricated by using the CIGS thin films deposited from these pastes showed
an improved efficiency (4.66%) over those fabricated using CIGS thin-film deposited from thicker
paste only (2.90%).
Deposition of CIGSSe absorber layers by electrospraying a propylene glycol solution of
copper(II), In(III), and Ga(III) nitrates as metal precursors has been reported (Yoon et al. 2014).
The bandgap energy of the absorber layer was found to be 1.15 eV. Solar cells with Mo/CIGSSe/
CdS/i-ZnO/n-ZnO/Ni/Al structure have been fabricated. The solar cell demonstrated a power
232 Solar Energy Conversion and Storage

conversion efficiency of 4.63% with an open-circuit voltage of 410 mV, a short-circuit current of


21 mA cm−2, and a fill factor of 0.5337 for an active area of 0.46 cm2.
Harvey et al. (2013) have reported the colloidal preparation of CIGS nanoparticles with a stoi-
chiometric composition of Cu0.8In0.7Ga0.3Se2. The nanocrystals were used as ink for the deposition
of thin films onto Mo-coated glass using the spray deposition process. More uniform and thicker
CIGSe films were obtained by multiple cycles of ink deposition and sintering. A power conversion
efficiency, as high as 7%, has been achieved by optimal use of this multiple deposition and sintering
approach.
Coughlan et al. (2013) have carried out an extensive study on the synthesis and shape evolution
of CuIn xGa1−xS2 nanocrystals and nanorods. These nanocrystals form ternary chalcopyrite CuInS2
nanoparticles, whereas incorporation of Ga into the lattice formed quaternary CIGS nanoparticles.
Nanorods with a controllable aspect ratio from 1.8 to 3.3 have been synthesized using these aliphatic
amines.
By varying the initial amount of metal precursors, a full range of stoichiometric compositions from
CuGaS2 to CuInS2, through intermediate CuInxGa1–xS2 with tunable x were synthesized. The bandgap
of the as-synthesized nanocrystals significantly varied with the composition, and an increase in band-
gap energy from 1.48 to 2.2 eV was observed with increasing gallium composition. Furthermore, the
average size of the nanoparticles could be varied from 13 to 19 nm by employing different reaction
temperatures ranging from 230°C to 290°C. A ligand exchange reaction carried out to replace organic
capping ligand with the inorganic S2− capping ligand resulted in only 50% exchange. Thin films pre-
pared by using these partially ligand exchanged nanocrystals exhibited high conductivity.
A solution-based approach has been used for the synthesis of wurtzite phase CuIn xGa1–xS2
nanocrystals with the value of x ranging from 0 to 1 (Wang et al. 2011). As-synthesized CuInS2
nanocrystals had a black color, whereas CuIn0.5Ga0.5S2 and CuGaS2 nanocrystals had dark red and
dark yellow colors, respectively. The TEM images revealed the formation of bullet-like CuInS2 nano-
crystals having a uniform size with 16 nm width and about 35 nm length. The morphology changed
to rod-like for the CIGS composition CuIn0.75Ga0.25S2 with a shorter length (Figure 11.1). In general,
the length of nanorods decreased with an increase in Ga concentration. When octadecene was used
as solvent, the morphology changed from nanobullets to nanorods, nanospheres, and nanotadpoles.
X-ray diffraction studies showed that the nanocrystals had a wurtzite crystallographic structure.
The bandgap of as-synthesized CuIn xGa1–xS2 nanocrystals increased with increasing Ga composi-
tion from 1.53 eV for CuInS2 to 2.48 eV for CuGaS2. Due to the relatively large size of nanocrystals,
no broadening of bandgap due to quantum confinement was observed.

11.3.4  Copper Indium Gallium Diselenide (CIGSe)


Tang et al. (2008) have reported the synthesis of monodispersed, high-quality CuGaSe2, CuInSe2,
and Cu(InGa)Se2 nanoparticles. It was demonstrated that the nanoparticle size can be tuned by
carefully choosing the reaction temperature, whereas the desired stoichiometric composition can
be attained by suitably adjusting the precursor concentrations. As-grown CuGaSe2 and CuInSe2
nanocrystals had bandgap energy of 1.68 and 1.01 eV, respectively.
Panthani et al. (2008) introduced the concept of using nanoparticles-based ink for fabrication
of an absorber layer in solar cells. They have reported the synthesis of chalcopyrite CuInS2 and
CuIn xGa1–xSe2 (CIGSe) by arrested precipitation in solution. As-synthesized CIGSe nanocrystals
had a mean diameter ranging from 5 to 25 nm as revealed by TEM analyses. It was demonstrated
that the relative atomic ratio (In/Ga) in the CIGS nanocrystals could be controlled by suitably vary-
ing the ratio of In/Ga precursors used in the reaction. The bandgap of CIGSe nanocrystals showed
significant variation with change in stoichiometric composition of the nanocrystals. Uniform, crack-
free films of micrometer thickness were fabricated by using as-prepared nanocrystal ink. Prototype
photovoltaic devices were constructed to demonstrate a proof of concept for a reproducible photo-
voltaic response.
Nanomaterials for Solar Energy 233

(a) (b) (c)

20 nm 20 nm 20 nm

(d) (e) (f )

>
02
<0
0.323 nm

<0
02 0.317 nm
> 0.322 nm

>
00
<1
5 nm 5 nm 5 nm

FIGURE 11.1  TEM images of (a) bullet-like CuInS2 nanocrystals synthesized in OLA, (b) rod-like
CuIn0.75Ga0.25S2 nanocrystals synthesized in ODE, and (c) tadpole-like CuGaS2 nanocrystals synthesized in
ODE; HRTEM images of individual (d) bullet-like, (e) rod-like, and (f) tadpole-like nanocrystals. (Adapted
from Wang, Y.-H. A., X. Zhang, N. Bao, B. Lin, and A. Gupta, J. Am. Chem. Soc., 133, 11072–11075, 2011.
With permission.)

Preparation of CuIn1–xGa x(S1–ySey)2 nanocrystals with controllable stoichiometric composition (x,


y ranging from 0 to 1) and graded bandgaps has been reported (Chang et al. 2011). Stoichiometric
composition of the CuIn1–xGa x(S1–ySey)2 nanocrystals was successfully controlled by using variable
ratios of In and Ga as well as S and Se reactants in the reaction. It was observed that tuning of
optical bandgap from 0.98 to 2.40 eV was possible by suitably varying the composition of CIGSSe
nanocrystals.
Jiang et al. (2012c) have used colloidal nanocrystals capped with metal chalcogenide complexes
as soluble precursors for CuInSe2, CIGSe, and CZTS materials for solar cell applications. TEM and
p-XRD studies demonstrated that the ligand exchange reaction proceeds without altering the phase
and morphology of Cu2–xSe nanocrystals. Thin films deposited from metal chalcogenide complex
of Cu2–xSe by spin coating or spray coating were subsequently annealed at 500°C, and solid-state
reactions resulted in complete transformation of In2Se42−-capped Cu2–xSe nanocrystals into CuInSe2
thin films without any cracks. The bandgap of the as-deposited CuInSe2 thin films was found to
be 1.01 eV. This approach has also been used to deposit high-quality semiconductor thin films of
CIGSe and CZTS.
A new versatile solution-based strategy for in situ preparation of metal–organic molecular pre-
cursors based on butyldithiocarbamic acid has been reported (Wang et al. 2012). Carbon disulfide
and 1-butylamine undergo a reaction producing butyldithiocarbamic acid, which reacts with a vari-
ety of metal oxides and hydroxides to form thermally degradable metal–organic precursors. This
approach has been used for the fabrication of a CIGSSe-based absorber layer in thin-film solar cells.
SEM images showed compact and dense film without any noticeable cracks. Solar cells fabricated
with the configuration glass/Mo/CIGSSe/CdS/i-ZnO/ITO/Al demonstrated a power conversion
efficiency of 8.8%. It was suggested that optimization of film thickness and bandgap can further
improve performance of the solar cells.
234 Solar Energy Conversion and Storage

Ahmadi et al. (2012) have carried out a systematic study to investigate the growth mechanism
and evolution pathways of quaternary CuIn0.5Ga0.5Se2 nanocrystals in a hot coordinating solvent. A
mixture of binary Cu2–xSe and ternary CuGaSe2 nanocrystals is formed. After 40 min of reaction,
a small amount of CIGSe was observed to form by reaction of Cu2–xSe, CuGaSe2 nanocrystals and
amorphous In2Se3 nuclei. At 60 min, a biphasic CuIn xGa1–xSe2 mixture with variable x values was
formed which underwent subsequent transformation to monophasic CuIn0.5Ga0.5Se2 after another
60 min.
Synthesis of CuIn0.7Ga0.3Se2 (CIGSe) nanopowders has been carried out by microwave irradia-
tion (Seelaboyina et al. 2013). CIGSe nanopowders thus synthesized had particle size ranging from
20 to 80 nm, whereas the bandgap was found to be 1.1 eV. CIGSe nanopowders were obtained by
centrifugation at 5000–10,000 rpm, which was dried at 100°C–200°C for 4–8 h. Finally, CIGSe ink
was prepared by ball milling the powder in isopropanol and 1,2-propanediol for 2–10 h. X-ray dif-
fraction studies showed that the CIGSe nanopowders had a chalcopyrite crystallographic structure
corresponding to standard ICDD pattern 035-1102 for CuIn0.7Ga0.3Se2. SEM, TEM, and dynamic
light scattering (DLS) results showed particle size to be less than 100 nm, whereas elemental map-
ping confirmed uniform distribution of constituent elements throughout the powder particles.
A process for nonvacuum deposition of CIGSe absorber layers from hydroxide-containing par-
ticles and subsequent selenization by Se vapors has been reported (Uhl et al. 2013). Another type
of CIGSe thin film was prepared by sequential deposition of multiple layers by varying precursor
inks (In + Ga/Cu/In + Ga) and repeating the drying and deposition steps. These films were also
selenized at 550°C. Any residual Cu–Se phases were removed by etching of the selenized films with
KCN. CIGSe absorber formed from deposition of the inks mixture was mainly porous; however,
sequential deposition of inks resulted in improved sintering properties. In situ XRD measurements
showed that CuSe2, CuSe, and Cu2–xSe are formed as binary phases at 220°C to 350°C in the stacked
film, while only Cu2–xSe was detected from 270°C to 440°C in the case of mixed precursor ink
deposition. Power conversion efficiencies up to 4.8% and 5.8% were achieved with the absorber
layer deposited from mixed inks and stacked deposition, respectively, despite the porosity and large
compositional gradients in thin films.
Synthesis of CuIn0.7Ga0.3Se2 nanocrystals by a facile heating-up process using metal chloride
salts, selenium, octadecene, and oleylamine has been reported (Hsu et al. 2013b). It was observed
that with the supersaturation, formation of CuSe nanocrystals occurs first, whereas amorphous
indium selenide and gallium selenide nuclei are later formed on the surfaces of CuSe nanocrystal-
lites at the lower temperature and shorter reaction time. Finally, the amorphous indium selenide and
gallium selenide nanocrystals react with CuSe crystallites at high temperature (~250°C), yielding
thermodynamically stable chalcopyrite CIGSe nanocrystallites.
Li et al. (2013c) have reported the synthesis of ternary and quaternary CuIn1–xGa xSe2 nanocrystals
by an organo-alkali–assisted diethylene–glycol solution-based approach. Monophasic, well-dis-
persed, and granular chalcopyrite nanocrystals with controlled stoichiometry (0 ≤ x ≤ 1) and a size
ranging from 10 to 20 nm were synthesized using the hot-injection approach. The best results were
obtained by using triethylenetetramine with the additional range of 2–4 vol% into a diethylene gly-
col–based solution. The In/Ga atomic ratios of the as-synthesized CuIn1–xGa xSe2 nanocrystals were
very similar to those in the feeding precursor solutions in the case of TETA-assisted diethylene gly-
col synthesis. Tuning of the optical bandgap from 1.05 to 1.7 eV was feasible by controlled variation
of the In/Ga ratio in CuIn1–xGa xSe2 nanocrystals.
A microwave-assisted solvothermal synthetic method has been used for the preparation of mul-
tiphase CIGSe nanoparticles (Seo et al. 2014). By using this approach, composition-controlled syn-
thesis of CIGSe materials with Cu to (In + Ga) atomic ratio varying from 0.6 to 1.0 has been carried
out, and dense absorber layers have been fabricated. Finally, a photovoltaic device based on an as-
deposited CIGSe absorber layer was fabricated, and its performance parameters were investigated.
A ligand exchange method has been used where readily synthesized CuInSe2 and CuIn1–xGa xSe2
(CIGSe) nanocrystals capped with organic oleylamine are converted to 1-ethyl-5-thiotetrazole–capped
Nanomaterials for Solar Energy 235

nanocrystals (Lauth et al. 2014). The ligand exchange reaction proceeds with no deleterious effect
on the properties of nanoparticles, and colloidal stability of the chalcopyrite materials is preserved
in the process. Measurement of the current–voltage characteristics of thus-obtained nanocrystal
films before and after thermolysis of ligand were carried out in the dark and under illumination.
It was observed that the conductivity of trigonal pyramidal CuInSe2 nanocrystals demonstrated an
increase by four orders of magnitude for ligand-free nanocrystal films. Similarly, a two orders of
magnitude improvement in the photoconductivity of the CIGS nanocrystal film was observed.
A one-pot solution-based approach has been used for the synthesis of monophasic, chalcopyrite
CuIn1–xGa xSe2 nanoparticles with controlled stoichiometric composition (0  ≤  x  ≤  1) (Han et al.
2014). The effects of the Ga/In + Ga atomic ratio were determined on crystallographic structure,
shape, and optical properties of the as-prepared CuIn1–xGa xSe2 nanoparticles. The SEM images
revealed polydispersed size and shape of the nanoparticles, and nanoparticles with higher Ga con-
tent showed a pronounced tendency toward agglomeration. Bandgap energy of the as-synthesized
nanocrystals exhibited an increase from 1.00 to 1.68 eV with increasing Ga/In + Ga atomic ratio.
Roux et al. (2013) have developed a new ink-based process for fabrication of an efficient chal-
copyrite CuIn1–xGa xSe2–based absorber for use in thin-film solar cells. Inks prepared by using
as-synthesized nanoparticles could be safely coated under ambient conditions to form thin films
by commercially available techniques like doctor blading. Annealing of the precursor thin films
under Se vapors in a primary vacuum formed a functional absorber layer for thin-film solar cells.
CuIn1–xGa xSe2 thin films formed by employing this simple, two-step process exhibited strong
mechanical adhesion. Solar cells fabricated using these absorbers demonstrated a power conversion
efficiency higher than 7%.

11.4  COPPER-ZINC-TIN CHALCOGENIDES (CZTSSe)


Considerable research attention has been given toward Cu2ZnSn(S,Se)4 (CZTSSe) system due to the
potential of these materials for the production of low-cost solar cells with good power conversion
efficiency (PCE). This material has a bandgap ranging from 1.0 to 1.5 eV, with high optical absorption
(>104 cm) and is based on low-cost, nontoxic, and Earth-abundant elements. Therefore, it has emerged
as a promising material that can lead to both economically and ecologically sustainable production
of solar cells without any issue of scarcity of constituent elements. PCE of CZTSSe-based thin films
has already reached 12% during the last few years. Metal chalcogenide–based materials can tolerate a
higher degree of either structural or electronic defects and, hence, can be deposited in a cost-effective
manner by employing non-vacuum-based deposition approaches. In the case of CZTSSe, non-vacuum-
based solution processing has provided device performances that are even better than those achieved
through vacuum-based approaches. Processing of the CZTSSe must be controlled to form a pure
kesterite phase only, and this can be achieved only within a limited stoichiometric range with [Cu/
(Zn + Sn)] = 0.8–0.9 and Zn/Sn = 1.1–1.4. Deviation from this stoichiometry leads to the formation
of binary phases (Cu, Sn, and Zn chalcogenides) and/or ternary phases (copper tin chalcogenides).
Due to the volatile nature of constituent elements and binary phases, phase control becomes some-
what difficult, especially in deposition techniques involving ultrahigh vacuum and high temperatures.
Furthermore, in solution-based approaches, phase formation occurs in the liquid phase, thus allow-
ing better diffusivity for constituting elements in order to overcome the barriers of activation energy
and form a thermodynamically stable phase. Therefore, solution-based approaches providing targeted
stoichiometric composition at relatively lower temperatures have demonstrated better control over the
formation of monophasic CZTSSe. Device performance is also dependent on defect properties of the
CZTSSe. A dependable synthetic strategy is vital for controlling the defect profile of the final photoab-
sorber film, and nanoink-based approaches provide a facile way of incorporating extrinsic dopants for
the optimization of defect structure and passivation of grain boundaries.
The first step in a CZTSSe nanocrystals–based approach is obviously the synthesis of high-
quality kesteritic nanocrystals. These nanocrystals along with additives are then deposited as thin
236 Solar Energy Conversion and Storage

films by a variety of commercially available methods like spin coating, doctor blading, inkjet print-
ing, and so on. The final step involves post-treatment procedures like annealing in a sulfur/selenium
environment. This section will mainly focus on various synthetic strategies for obtaining high-
quality CZTSSe nanocrystals for use as inks in solar devices.
The first-ever synthesis of Cu2ZnSnS4 nanocrystals and their use in the fabrication of solar cells
was reported by Guo et al. (2009). The resulting CZTS nanocrystals had sizes ranging from 15 to
25 nm and a bandgap of 1.5 eV. These nanocrystals were drop coated onto Mo-coated soda lime
glass substrates and then selenized at 500°C to form Cu2ZnSnSySe1–y (CZTSSe). Photovoltaic devices
fabricated using these CZTSSe films demonstrated a power conversion efficiency of 0.74% under
AM 1.5 illumination. Riha et al. (2009) carried out synthesis of homogeneous and nearly mono-
dispersed CZTS nanocrystals by hot-injection approach. The mean diameter of the thus-prepared
nanocrystals was found to be 12.8 ± 1.2 nm, and they had a bandgap of 1.5 eV. Using the same
method, Guo et al. (2010) synthesized CZTS nanocrystals with copper-poor and zinc-rich stoichi-
ometry. Hexanthiol-dispersed nanocrystals were deposited as 1 µm–thick film by knife coating onto
Mo-coated soda lime glass. The films after post-treatment processes had a Cu/(Zn + Sn) = 0.79 and
Zn/Sn = 1.11 composition. Solar cells fabricated by using these films showed 7.2% power conver-
sion efficiency under one sun illumination.
Riha et al. (2011) extended their work to the compositionally controlled synthesis of phase-pure
Cu2ZnSn(S1–xSex)4 nanocrystals (with 0 ≤ x ≤ 1). Compositional control of the chalcogenide ratios
was obtained by sonicating the targeted amounts of S and Se along with sodium borohydride and
oleylamine to balance S and Se reactivities. A probe into lattice parameters and bandgap energies
of the composition tunable CZTSSe was carried out to investigate the effect of Se on material prop-
erties of finally deposited film. With an increase in the amount of Se, lattice parameters showed
an increase, bandgap was slightly decreased, and the electrical conductivity of the nanocrystals
increased due to more pronounced grain growth and passivation of grain boundaries.
A facile colloidal synthesis of CZTS nanocrystals was carried out using metal dithiocarbamic
acid salt [Cu2ZnSn(S2CNEt2)10] as the precursor, oleylamine as the activation agent, and oleic acid
as the capping agent, respectively (Zou et al. 2011). EDX analyses showed that the CZTS nanocrys-
tals had a sulfur-poor stoichiometric composition; ratio of Cu:Zn:Sn:S being 26:14:18:42 precisely.
Oleylamine balanced the reactivity of cationic constituents, while oleic acid bound to the surface of
growing CZTS nanocrystals.
Guo et al. (2012) also prepared CZTGeSSe nanocrystals by partial substitution of tin precursor
with germanium precursor. The bandgap energy of the CZTGeSSe nanocrystals was found to be
1.09 eV. The device fabricated using these nanocrystals as the absorber layer demonstrated 8.4%
PCE without the use of any antireflection coating.
Synthesis of CZTS nanocrystals with the average nanocrystal diameter ranging from 2 to 7 nm
was carried out by heating a stoichiometric mixture of copper, zinc, and tin diethyl dithiocarbamate
complexes (Khare et al. 2011). It was observed that the size of the nanocrystals could be tuned by
varying the growth temperature and the amount of oleylamine used, whereas the reaction dura-
tion had no effect on the size of the nanocrystals. Nanocrystals with average diameters of 2.0 and
2.5 nm showed a significant shift in their optical absorption spectra. Similarly, colloidal prepara-
tion of CZTS nanocrystals using copper ethylxanthate, zinc ethylxanthate, and tin(IV) chloride
has been reported by Liu et al. (2011). CZTS nanocrystals had a mean diameter of 15.6 ± 2.0 nm
and showed a tendency to grow larger with increased reaction durations. Stoichiometric composi-
tion of the nanocrystals was determined by EDX analyses, which showed Cu:Zn:Sn:S ratios to be
1.81:1.17:0.95:4.07. A bandgap of the nanocrystals estimated from their optical absorption spectra
was found to be 1.5 eV.
Good-quality monodispersed CZTS nanoparticles having thermodynamically stable kesterite
and wurtzite phases have been synthesized via a simple, one-pot, low-cost solution method (Cattley
et al. 2013). The nanoparticles prepared at 140°C had the tetragonal CZTS crystallographic struc-
ture, whereas those grown at higher temperature had a wurtzite crystallographic phase as revealed
Nanomaterials for Solar Energy 237

by high-resolution transmission electron microscopy (HRTEM), selected area electron diffraction


(SAED), and XRD analyses.
Spindle-shaped CZTS nanoparticles of the kesterite phase have been prepared by hot-injection
approach using corresponding metal chlorides as metal precursors and thiourea as the sulfur source
(Wei et al. 2012). The nanoparticles had an average length of 22.5 ± 2.0 nm and average width of
13.9 ± 1.5 nm. The bandgap of the nanocrystals was found to be 1.54 eV. Formation of spindle-like
morphology and narrow size distribution was attributed to the gradual release of sulfur from thio-
urea into solution, resulting in homogeneous nucleation and growth.
A hot-injection approach involving the injection of concentrated metal precursors (copper(II)
acetylacetonate, zinc acetate, and tin(IV) acetate) and elemental sulfur solution in oleylamine
into preheated triphenylphosphate (TPP) as capping ligand has been reported (Kim et al. 2013). It
was observed that the TPP ligand surrounding the CZTS nanocrystals decomposes more readily
than conventionally employed capping ligands such as oleylamine, thus giving a better absorber
layer for solar cell applications. As-synthesized CZTS nanocrystals had an average diameter of
14.5  ±  4.6  nm, whereas the bandgap was calculated as 1.5  eV. Photovoltaic performance of the
CZTS absorber layer was determined by fabricating a solar device with a configuration Mo-coated
SLG/CZTS/CdS/i-ZnO/indium tin oxide (ITO)/patterned Ni/Al grid. The device showed a power
conversion efficiency of 3.6% without any Se treatment.
Cao et al. (2013) have reported the synthesis of CZTSSe nanocrystals through hot-injection
approach using copper(II) acetate, zinc acetylacetonate, tin(IV) acetate, elemental sulfur, selenium,
and oleylamine. Cu2ZnSn(S1−xSex)4 nanocrystals (x = 0, 0.2, 0.5, 0.85, and 1.0) were prepared and
used as counterelectrodes in quantum dot–sensitized solar cells (QDSSCs). It was observed that all
of the QDSSCs using CZTSSe as counterelectrodes demonstrated superior current–voltage charac-
teristics than the Pt-based QDSSCs, regardless of the selenium amount in CZTSSe. The QDSSC
with CZTSSe (x = 0.5) counterelectrode showed the highest energy conversion efficiency of 3.01%,
which was higher than that (1.24%) obtained using platinum counterelectrode.
The CZTS nanocrystals have been prepared by the hot-injection method using correspond-
ing diethyldithiocarbamate complexes of copper, zinc, and tin in oleylamine at 280°C for 10 min
(Chernomordik et al. 2014a). Thin films based on nanocrystals were prepared by drop casting the
colloidal CZTS suspension onto soda lime glass and quartz substrates. The effect of various param-
eters like annealing temperature, annealing time, and sulfur vapor pressure on the evolution of
microstructure of CZTS nanocrystal films was thoroughly investigated.
Guan et al. (2014a) have reported the synthesis of CZTS nanocrystals by a hot-injection method.
It was observed that crystalline CZTS and Cu2SnS3 and amorphous ZnS were formed after injec-
tion of sulfur–oleylamine into the mixture solution. The reaction of ZnS with Cu2SnS3 forms CZTS
during sulfur annealing at a high temperature. Therefore, in case of a much higher Zn/Sn ratio, the
excess ZnS cannot be eliminated and appears as an impurity phase. When the Zn/Sn precursor ratio
was reduced to 0.6, a relatively pure CZTS phase was obtained.
Synthesis of CZTSe nanoparticles by the hot-injection approach using bis-(triethylsilyl)selenide
[(Et3Si)2Se] as the selenium source has been reported (Jin et al. 2014). Near stoichiometric and
monophasic nanocrystals with average size ranging from 25 to 30 nm were prepared. The presence
of secondary and ternary phases in the material was ruled out by XPS and Raman spectroscopy
measurements. The prepared CZTSe nanocrystals had an optical bandgap of 1.59 eV.
Steinhagen et al. (2009) have synthesized the CZTS nanocrystals through a high-temperature
arrested precipitation approach using oleylamine (OLA) as the coordinating solvent. EDX showed
average composition of the as-prepared nanocrystals to be Cu2.08Zn1.01Sn1.20S3.70, and average diame-
ter of the crystals was found to be 10.6 ± 2.9 nm. The bandgap of the nanocrystals was calculated to
be 1.3 eV. Solar PV devices were fabricated by deposition of CZTS thin films through spray coating
the toluene dispersion of nanocrystals. No annealing or other post-treatment steps were performed.
Thus, fabricated devices with Au/CZTS/CdS/ZnO/indium tin oxide (ITO) configuration showed
0.23% PCE under AM 1.5 illumination.
238 Solar Energy Conversion and Storage

Thermal reactions of metal acetate and elemental sulfur in hot oleylamine were used to prepare
CZTS nanoparticles (Kameyama et al. 2010). It was observed that the nanocrystals deposited at
240°C or higher temperature have pure kesterite phase, whereas those deposited at 180°C or below
showed broad peaks originating from the CuS phase. CZTS nanoparticles had an average size of
5.6 nm, and bandgap of the nanocrystals was determined to be 1.5 eV.
Due to improved device quality by partial replacement of indium in CuInSe2 with a lower atomic
number gallium to form CIGSe, tin atoms in CZTS were partially replaced by smaller group IV
element germanium to give Cu2Zn(Sn1–xGex)S4 (Ford et al. 2011). Alloying germanium in CZTS
nanocrystals opened up a new window for optimizing the bandgap by suitably adjusting the Ge/
(Sn + Ge) ratio. CZTGSSe nanocrystals thus synthesized were polydispersed ranging in size from 5
to 30 nm (Figure 11.2). The bandgap of the as-synthesized nanocrystals was found to increase with
increasing Ge content in the CZTGSSe lattice. The solar cells with SLG/Mo/CZGSSe/CdS/i-ZnO/
ITO/Ni-Al architecture showed a power conversion efficiency 6.8% for CZTGS nanocrystals synthe-
sized with Ge/(Ge + Sn), Cu/(Zn + Sn + Ge), and Zn/(Sn + Ge) ratios of 0.7, 0.8, and 1.2, respectively.
Phase-pure wurtzite CZTS nanocrystals were synthesized by Li et al. (2012a). It was found that
the use of dodecanethiol alone in the reaction leads to the formation of coexisting wurtzite and
kesterite phases. However, when some amount of oleylamine was added, the reaction environment
was changed, and only the wurtzite phase was obtained. Rice-like CZTS nanocrystals had a mean
diameter ranging from 10 to 40 nm depending on the reaction duration. The preparation of wurtzite
CZTS nanocrystals based on arrested precipitation has been reported, which utilizes acetates of
copper, zinc, and tin as a metal source—diethanolamine as solvent and thiourea as sulfur source
(Li et al. 2012b). It was reported that the initial reaction at 160°C and the amount of thiourea are key
factors controlling the selective growth of CZTS nanocrystals as the wurtzite phase. Nanocrystals
thus formed had a mean diameter of 10 ± 1.1 nm with a bandgap of 1.56 eV.
The synthesis of CZTS and CZTSe nanoparticles has also been carried out by using metal halide
salts CuI, ZnCl2, and SnI4 and sulfur/selenium (Rath et al. 2012). CZTS nanocrystals grown at

(a) (b)

25 nm 25 nm

(c) (d)

5 nm
25 nm

FIGURE 11.2  TEM images of (a) CZTS, (b) CZTGS, and (c) CZGS. (d) HR-TEM of CZGS nanocrystals.
(Adapted from Ford, G. M., Q. Guo, R. Agrawal, and H. W. Hillhouse, Chem. Mater., 23, 2626–2629, 2011.
With permission.)
Nanomaterials for Solar Energy 239

210°C had an average size ranging from 7 to 10 nm. The EDX analysis showed near stoichiometric
composition for CZTS nanoparticles, whereas the Se analogue prepared using the same methodol-
ogy had no zinc and was found to be Cu2SnSe3. Yang et al. (2012) have reported the synthesis of
CZTS nanocrystals and their utilization as a potential thermoelectric material. The as-prepared
CZTS nanocrystals had a mean diameter of 10.6 ± 1.9 nm, and bandgap was found to be 1.51 eV.
The nanocrystals were then compressed into compact pellets by spark plasma sintering (SPS) and
hot press. A significant enhancement in Seebeck coefficient and decrease in thermal conductivity in
comparison with bulk crystals was observed in electrical and thermal measurements between 300
and 700 K. It was also observed that doping of CZTS nanocrystals with extra copper significantly
increased the electrical conductivity and decreased the thermal conductivity.
Chesman et al. (2013a) have reported a gram-scale noninjection synthesis of CZTS nanocrys-
tals. Nanocrystals were grown using CuI, SnCl4.5H2O, zinc ethyl xanthate, and dodecanethiol in
addition to oleylamine at 250°C for 30 min. It was observed that due to different decomposition
temperatures of sulfur precursors. Morphology of the nanocrystals was predominantly triangular,
and the mean size was in the range of 5.5–7.8 nm. Optical bandgap of CZTS nanocrystals ranged
from 1.45 to 1.55 eV. This one-pot, noninjection synthetic strategy was extended by using CS2 to
form oleyldithiocarbamate and dodecyltrithiocarbonate in situ as the sulfur source (Chesman et al.
2013b). It was observed that the size of the nanocrystals was 1.6 ± 0.4 nm at 138°C and it grew to
5.0 ± 0.1 in 10 min period after decomposition of dodecanethiol at 250°C. After 30 min reaction at
250°C, the average size of the crystallites was recorded as 7.4 ± 0.1 nm.
Liao et al. (2013) have reported a simple, one-pot, noninjection strategy for large-scale synthesis
of wurtzite CZTS nanocrystals. The bandgap of the CZTS nanocrystals was 1.5 eV, while the aver-
age size was 13 nm. The nanocrystals had somewhat bullet- or leaf-like morphology. High-quality
w-CZTS nanocrystals demonstrated high crystallinity, a monophasic nature, a uniform composi-
tion, and narrow size distribution.
A phosphine-free approach for the synthesis of high-quality CZTSe nanocrystals in organic
solvents has been reported (Liu et al. 2012).
Wurtzite CZTS nanorods have been synthesized using commercially available cationic pre-
cursors (Thompson et al. 2013). Clear variation in the composition of rods along their axes was
observed, and a Cu-rich end and a Zn-rich end could be identified. The copper precursor had the
most reactivity and the highest rate of nucleation followed by the zinc precursor, while tin had the
least reactivity. The aspect ratio of the CZTS was increased by decreasing the initial loading of
the most reactive Cu precursor. In this case, three metal precursors nucleate at more comparable
rates; thus, longer nanorods with more homogeneous elemental composition along their axes were
formed. Increasing the initial loading of Cu precursor results in a low aspect ratio (formation of
dots) and/or phase segregation (yielding the binary phases).
Synthesis of CZTS nanocrystals by thermolysis of single-source precursors of copper, zinc,
and tin diethyldithiocarbamates has been reported (Chernomordik et al. 2014b). CZTS nanocrys-
tals having mean size between 2 and 40 nm were synthesized by varying the growth temperature
between 150°C and 340°C. The kesterite phase of the nanocrystals was confirmed by x-ray dif-
fraction, Raman spectroscopy, and transmission electron microscopy, whereas energy dispersive
x-ray spectroscopy confirmed stoichiometric composition of the CZTS nanocrystals. Nanocrystals
having 2 nm size had a bandgap of 1.67 eV exhibiting strong quantum confinement, whereas larger
nanocrystals had a bandgap of 1.5 eV.
Todorov et al. (2010) used a hydrazine-based nonvacuum, slurry-coating method. The slurry (or
ink) used for deposition of CZTSSe was composed of a Cu–Sn chalcogenide (S or S–Se) solution
in hydrazine and particle-based Zn-chalcogenide precursors, ZnSe(N2H4) or ZnS(N2H4) formed in
situ upon the addition of Zn powder. The CZTSSe obtained had a zinc-rich and copper-poor stoi-
chiometric composition with ratios Cu/(Zn/Sn) = 0.8 and Zn/Sn = 1.22. This material was deposited
onto Mo-coated glass for fabrication of a solar device, which demonstrated a PCE of 9.6%. A simi-
lar technique was used to prepare 2.5 µm thick CZTSSe thin films with S/(S + Se) ratio of 0.4 ± 0.1
240 Solar Energy Conversion and Storage

(Barkhouse et al. 2012). Deposition of CdS buffer, a ZnO window layer, and indium tin oxide (ITO)
was carried out by using CBD and RF magnetron sputtering. An antireflection MgF2 coating was
deposited using electron beam evaporation. The finished device demonstrated a power conversion
efficiency of 10.1% under AM 1.5 illumination conditions. CZTSe-based thin-film solar absorbers
with some percentage of Sn atoms replaced by smaller isoelectronic Ge atoms were also prepared.
These films were prepared using the hydrazine-based mixed particle-solution approach pioneered
by this group. They observed an increase in bandgap of the absorber layer from 1.08 to 1.15 eV with
40% substitution of Sn atoms with Ge in CZTSe thin film. Devices fabricated by using this absorber
layer showed a power conversion efficiency of 9.1% and a high open-circuit voltage.
An optical-design approach has been reported, which resulted in improvement of the
short-circuit current and power conversion efficiency of CZTSSe solar cells (Winkler et al. 2014).
They have optimized the thickness of the upper-device layers in a way that allows maximum trans-
mission into the CZTSSe absorber. This design approach was based on optical modeling of ideal-
ized planar devices with a semi-empirical approach for treating the impact of surface roughness.
Effectiveness of the new device architecture based on thinner CdS and TCO layers was experi-
mentally demonstrated by fabricating a solar device with an overall power conversion efficiency of
12.0%—a new record in CZTSSe-based devices.
Solvothermal synthesis of spherical CZTS nanoparticles using ethylene glycol, copper (II) chlo-
ride dehydrate, zinc (II) chloride and tin (IV) chloride tetrahydrate, and thiourea and PVP was
reported by Zhou et al. (2011a). Nanocrystals of about 100–150  nm size agglomerated to form
spheres of kesteritic CZTS having bandgap of 1.48 eV. The same group has also reported the prep-
aration of hierarchal flower-like CZTS nanostructures through a similar solvothermal approach
(Zhou et al. 2011b). The effect of various synthesis parameters such as reaction temperature, reac-
tion duration, precursor concentration, and amount of PVP on size and morphology of the final
CZTS nanostructures was investigated. Formation of flower-like morphology was attributed to a
nucleation-dissolution–recrystallization mechanism during crystal growth.
The synthesis of surfactant-free, solvent-redispersible CZTS nanoparticles based on high-
temperature polycondensation reactions has been reported (Zaberca et al. 2012). Thiourea acted
both as complexing agent to inhibit growth of the nanoparticles and also as surfactant to aid redis-
persion of nanocrystals in polar solvents. Films of the CZTS absorber were formed on Mo-coated
glass by dip coating from a concentrated dispersion of nanocrystals with intermediate heat treat-
ment at 400°C. Crack- and defect-free sintered films formed from surfactant-free CZTS can be a
promising option for low-cost solar cells.
Arul et al. (2013) have reported the solvothermal synthesis of CZTS nanospheres with an average
diameter of 3.26 nm. Small CZTS nanospheres exhibited a bandgap of 1.84 eV, significantly larger
than the bandgap of bulk CZTS due to strong quantum confinement effects. An organic photovoltaic
cell fabricated by using these CZTS nanospheres showed a PCE of 0.952%. Mali et al. (2013) have
reported a simple, single-step synthesis of CZTS nanocrystals from corresponding metal halides,
thiourea, and PVP. Nanoparticles thus obtained had a tetragonal crystallographic structure with size
ranging from 4 to 7 nm. Synthesis of less than 10 nm CZTS nanocrystals by solvothermal reaction
at 180°C for 15 h has also been reported by Pal et al. (2013).
The CZTSe nanoparticles with a mean diameter of 200−300 nm have been prepared using a one-
step reaction without using any surfactant or template (Du et al. 2012b). The CZTSe nanoparticles
were drop-casted onto FTO substrate and used as counterelectrode (CE) in dye-sensitized solar cells
(DSSCs), which demonstrated an efficiency of 3.85%. A one-step, solvothermal treatment–based
approach has been used for in situ deposition of CZTS thin films onto stainless steel and FTO glass
substrates (Zhai et al. 2014). Characterization of the resulting material has been done by p-XRD,
SEM, TEM, UV-vis spectroscopy, and Raman spectroscopy. The effects of temperature, reaction
duration, and ratio of Cu/Zn/Sn precursors on the formation of CZTS nanocrystal films have been
investigated. Phase-pure CZTS thin films were obtained at a temperature of 250°C or higher. The
use of flexible substrates for deposition of CZTS thin films is attractive for industrial applications.
Nanomaterials for Solar Energy 241

Jiang et al. (2012d) have reported a metastable orthorhombic phase of CZTS prepared through
a hydrothermal approach using SnCl2, H2O, ZnCl2, CuCl2.2H2O, and thiocarbamide in a water-
ethylenediamine mixture at 200°C. Nanocrystals thus obtained had a crystal structure based on a
double-wurtzite cell having space group Pmn21. The bandgap of the orthorhombic CZTS nanocrys-
tal was 1.45 eV. Annealing of the nanocrystals at 500°C resulted in phase transformation, yielding
the tetragonal kesterite phase of CZTS.
CZTS nanoparticles powder has been prepared by using corresponding metal halides and
Na2S.9H2O (Mariama et al. 2013). An ethylene glycol suspension of milled CZTS nanoparticles
was deposited onto 1 mm FTO coated glass by spin coating. The bandgaps of thin films deposited
after 18, 12, and 6 cycles of spin coating were found to be 1.42, 1.57, and 1.67  eV, respectively.
Photoelectrical characterization showed a short-circuit current of 23.8 mA cm–2, open-circuit volt-
age as 0.394 V, fill factor as 28.78%, and efficiency as ~2.25%. Synthesis of phase-pure kesterite
CZTS nanocrystals with controllable size between 3 and 10.5 nm has been carried out using inor-
ganic metal salts copper (II) acetate monohydrate, zinc (II) nitrate, and tin (II) chloride and thiourea
by using the hydrothermal method (Liu et al. 2013a). The average size of the nanocrystals with
the reaction duration of 6, 12, 24, and 48 h at 180°C were 3.0, 4.8, 6.5, and 10.5 nm, respectively.
Bandgaps for CZTS nanocrystals with an average size of 10.5, 6.5, 4.8, and 3 nm were found to
be 1.48, 1.52, 1.61, and 1.89 eV, respectively. A slight blue shift in bandgap of the 4.8 nm CZTS
nanocrystals was observed, while a significantly large blue shift was observed for the 3 nm CZTS
nanocrystals due to the quantum confinement effect. The same group has also reported the hydro-
thermal synthesis of ~5 nm CZTS nanocrystals with an optical bandgap of 1.47 eV (Liu et al. 2013b).
Synthesis of good-quality CZTS nanocrystals by using a water-based one-step hydrothermal
method has been reported, and the influence of the sulfur precursor used in a hydrothermal reac-
tion has been investigated (Tiong et al. 2014). It was observed that the sulfur precursor significantly
affects the phase purity and crystal structure of the CZTS material. The use of organic sulfur
precursors like thioacetamide and thiourea gave a mixture of kesterite phase and wurtzite crystal
phase. However, pure kesterite CZTS was obtained by using Na2S as the sulfur source.
The CZTS nanocrystals–based ink has been prepared using a microwave-assisted synthesis
(Flynn et al. 2012). This method is capable of producing CZTS nanoparticles at significantly lower
temperatures and in lesser reaction duration. The CZTS nanocrystals were purified by removing the
unreacted species through centrifugation from de-ionized water and ethanol alternatively. The TEM
images showed polydispersed spherical particles having an average diameter of 7.2 ± 1.2 nm; SEM
images showed spherical clusters formed by the agglomeration of smaller nanoparticles. The band-
gap of the as-prepared nanoparticles was found to be 1.5 eV. Analyses using p-XRD and Raman
spectrometry confirmed that the CZTS nanocrystals correspond with the kesterite phase. Solar cells
fabricated by using these crystals showed a maximum power conversion efficiency of 0.25%. Shin
et al. (2012) have also reported the synthesis of CZTS nanocrystals through irradiation of aqueous
solutions of copper acetate, zinc acetate, tin chloride, and thioacetamide by microwave energy of
700 W for 10 min.
A rapid and facile deposition of CZTS thin films from a homogeneous precursor solution
directly onto conductive films has been carried out via thermolysis of precursors by microwave
heating (Knutson et al. 2014). The conductive films strongly absorb microwave energy, which
causes rapid heating to a sufficiently high temperature for the decomposition of precursors and
deposition of CZTS on the conductive layer. By using this approach, thin films having thicknesses
of 1–3 mm can be deposited in a very short time. Microwave annealing of the as-deposited CZTS
films in solvent resulted in an increase of the crystallite size, while no formation of impurity
phases was observed.
The synthesis of kesteritic CZTS nanocrystals using a continuous-flow mesofluidic reactor has
been reported (Flynn et al. 2013). Ethylene glycol was used as the solvent, and the composition of
the CZTS nanocrystals was controlled by judiciously varying the parameters, such as precursors
concentrations, reaction temperature, and residence time. The growth of CZTS nanocrystals from
242 Solar Energy Conversion and Storage

binary metal sulfide precursors was studied in detail. The average size of the nanocrystals obtained
was 5.4 ± 2.0 nm.

11.5  COPPER IRON TIN SULFIDE AND RELATED MATERIALS


The crystal structure of copper iron tin sulfide, Cu2FeSnS4 (CFTS), was first reported by Brockway
(1934). Bernardini et al. (1990, 2000) have investigated the crystallographic structure and mag-
netic properties of bulk single-crystalline CFTS. Synthesis of CFTS nanoparticles was reported
by hydrothermal reaction (Gui et al. 2004). The CFTS nanoparticles were formed as black pre-
cipitate. Studies using p-XRD showed the monophasic nature of the stannite material, and TEM
images revealed that the size of crystallites ranged from 10–20 nm. It was observed that formation
of a white colloidal precursor is important for subsequent conversion to the final multinary sulfide
(CFTS).
Zhang et al. (2012) have reported the synthesis of wurtzite CFTS nanocrystals by hot-injection
approach. The CFTS nanocrystals thus obtained had a mean size of approximately 20 nm. The x-ray
diffraction pattern of the nanocrystals synthesized at 210°C was similar to that of the simulated
pattern for the wurtzite phase of CFTS, whereas those grown at higher temperatures exhibited zinc
blende structure.
The stoichiometric composition of the CFTS nanocrystals grown at 210°C was slightly iron
deficient; however, it was improved by increasing the reaction temperature. The bandgap of the
nanocrystals grown at 210°C, 240°C, 270°C, and 310°C was found to be 1.54, 1.53, 1.51, and
1.49 eV, respectively.
Zhang et al. (2013b) have reported the preparation of wurtzite Cu2CoSnS4 via a simple thermal
decomposition approach. Nanocrystals thus formed had a nanorod-like morphology with average
length of 32 ± 2.0 nm and average width of 16 ± 1.5 nm. It was demonstrated that the stable stannite
crystallographic phase of Cu2CoSnS4 nanocrystals can be obtained by using higher reaction tem-
peratures or by postsynthesis annealing at above 400°C. Wurtzite nanocrystals exhibited a bandgap
of 1.58 eV.
Photoelectrical characterization of thin films based on Cu2CoSnS4 nanocrystals showed a pho-
toresponse indicating their potential for use in thin-film solar cells. A flexible chemical approach
for preparation of colloidal CFTS nanocrystals by thermal reaction of copper (II) acetate, iron (II)
chloride, tin (II) chloride, and sulfur in oleylamine has been reported (Li et al. 2012c). The p-XRD
pattern of the as-synthesized CFTS nanocrystals corresponded with the stannite CFTS. The aver-
age size of the as-synthesized nanocrystals, as revealed by TEM images, was 15–25  nm, which
increased to 20–40 nm after annealing. The bandgap was found to be 1.33 eV.
The synthesis of zinc blende CFTS nanocrystals by hot injection has also been reported by Liang
et al. (2012a). The p-XRD pattern of the as-prepared nanocrystals corresponded with the simu-
lated pattern for zinc blende CFTS with space group F-43 m and unit cell parameter a = 5.429 Å.
Nanocrystals had a spherical morphology with an average diameter of 7.5 nm. Magnetic studies by
superconducting quantum interference device (SQUID) revealed that the CFTS nanocrystals thus
obtained had ferromagnetic behavior, and bandgap was measured to be 1.1 eV. Liang and Wei (2012)
have also reported the synthesis of zinc blende CFTS nanocrystals using the same reaction strategy;
however, the nanocrystals had a lower bandgap of 0.92 eV.
A microwave irradiation–based method for the preparation of quaternary CFTS nanotubes with
an outer diameter of 400–800 nm and a thickness of 100–200 nm from corresponding metal halide
precursors has been reported (Lunhong and Jing 2012). Benzyl alcohol was used as the micro-
wave-absorbing solvent in this nonaqueous reaction. As-prepared CFTS nanotubes exhibited strong
absorption in the visible region and had a bandgap of 1.71 eV. Ai and Jiang (2012) have also reported
the synthesis of hierarchal porous quaternary CFTS hollow microspheres by a nonaqueous micro-
wave reaction. These microstructures showed excellent adsorption capacity. Furthermore, these
CFTS hollow microstructures demonstrated superb ability to remove organic pollutant from water.
Nanomaterials for Solar Energy 243

The synthesis of CFTS nanocrystals by hot-injection approach has been reported by Yan et al.
(2012). CFTS nanocrystals grown for 1 h had a triangular and spherical morphology with a mean
diameter of 13.2 ± 1.1 nm. Raman spectroscopy and p-XRD confirmed phase-pure deposition of
CFTS material, as no peaks assignable to Cu2SnS3 and FeS were observed. The EDX measurements
revealed that the CFTS nanocrystals had a sulfur-rich stoichiometry Cu2Fe1.05Sn1.06S4.69. The band
gap was found to be 1.28 ± 0.02 eV. The CFTS nanocrystals thus obtained were coated directly onto
ITO substrates and annealed at 350°C to form thin-film electrodes. Photoelectrochemical charac-
terization revealed p-type conductivity of the CFTS films with clear photoelectrochemical response
and good photostability. Values of the photocurrent density obtained for this system ranged from 11
to 13 mA cm–2 without any optimization.
The formation of highly crystalline and phase-pure quaternary chalcogenide (Cu2CoZnS4,
Cu2ZnSnS4) particles by a high-temperature route in molten KSCN at 400°C has been reported
(Benchikri et al. 2012). Metal oxides like Cu2O, SnO2, CoO, and ZnO were used as metal precursors.
It was demonstrated that the chemical homogeneity of various oxide precursors Cu2O–MO–(M =
Zn or Co)–SnO2–KSCN played a crucial role in the formation of a pure quaternary chalcogenide
structure. The nucleation rate of the quaternary chalcogenide particles was also affected remarkably
by the chemical homogeneity. These results thus highlight the ability to synthesize highly pure and
highly crystallized CZTS particles with controlled primary crystallite sizes.
A general strategy for the synthesis of quaternary semiconductor nanocrystals of Cu2CoSnS4,
Cu2FeSnS4, Cu2NiSnS4, and Cu2MnSnS4 based on solvothermal reaction has been reported by Cui
et al. (2012a). Cu2CoSnS4 and Cu2FeSnS4 nanocrystals possessed a zinc blende crystallographic
structure, whereas Cu2NiSnS4, and Cu2MnSnS4 had a wurtzite structure as revealed by p-XRD
measurements. The TEM images demonstrate that the Cu2CoSnS4 and Cu2FeSnS4 nanocrystals had
somewhat spherical morphology with a mean diameter of 5.6 ± 1.4 nm and 7.5 ± 1.5 nm, respec-
tively. Cu2NiSnS4, and Cu2MnSnS4 nanocrystals exhibited a nail-like and rod-like morphology.
Cu2FeSnS4, Cu2CoSnS4, Cu2MnSnS4, andCu2NiSnS4 nanocrystals possessed optical bandgaps of
1.19, 1.25, 1.28, and 1.49  eV, respectively. Photoelectrical measurements of the Cu2MnSnS4 thin
films prepared by spin coating a concentrated solution of Cu2MnSnS4 nanorod exhibited a photore-
sponse behavior, thus showing that these Cu2MnSnS4 nanorods can potentially be used for fabrica-
tion of low-cost PV solar cells.
Synthesis of Cu2CoSnS4 nanoparticles through low temperature, simple, and one-pot hydrother-
mal process has been reported (Murali and Krupanidhi 2013). The p-XRD studies showed that the
nanoparticles had a wurtzite crystallographic phase. A size ranging from 20 to 60 nm was observed
for Cu2CoSnS4 nanoparticles prepared under different conditions of reaction temperature, reaction
time, and surfactant. An unusual red-edge effect was exhibited by CCTS nanoparticles in all the
samples. Hybrid devices fabricated using these nanocrystals showed their potential use in solar
cells, photodetectors, and light-emitting diodes.
Park et al. (2013) have synthesized Cu2FeSnS4 nanoparticles with zinc blende crystal structure.
The TEM images revealed that the CFTS nanocrystals are well dispersed, having spherical mor-
phology. Average size of the nanocrystals varied from 11 to 15 nm, whereas the nanocrystals grew
to a size of 30–40 nm after annealing at 400°C for 15 min. Counterelectrodes for DSSC were fab-
ricated by spin coating the CFTS nanocrystals on FTO glass at 1000 rpm for 20 s, and the desired
thickness of the film was obtained after seven cycles of spin coating. It was demonstrated that the
DSSC based on CFTS as counterelectrode had efficiency comparable to that of DSSCs based on
Pt-based counterelectrode.
The Cu2(FexZn1–x)SnS4 nanocrystals with x ranging from 0 to 1 have been synthesized by using
the hot-injection approach in order to tune the bandgap of Cu2ZnSnS4 nanocrystals (Huang et al.
2013). Nearly monodispersed Cu2(FexZn1–x)SnS4 nanocrystals with sizes ranging between 10 and
20 nm were shown by TEM images. Raman spectroscopy and p-XRD studies demonstrated that
the Cu2ZnSnS4 and Cu2FeSnS4 crystallize as kesterite and stannite phases, respectively, and no
binary or ternary phases were formed. Shifting of the Raman peaks toward higher frequencies
244 Solar Energy Conversion and Storage

with increasing Zn content was also reported. It was observed that the bandgap of the nanocrystals
can be tailored from 1.25 eV for x = 0 to 1.5 eV for x = 1 by adjusting the composition variable x.
Improvement in the photoelectrical conversion performance was also observed when x = 0.4.
Synthesis of CFTS nanoparticles by the solvothermal route using N,N-dimethylformamide
(DMF) as solvent has been reported (Jiang et al. 2013). Peaks in the x-ray diffraction pattern of
as-synthesized nanocrystals could be indexed to the stannite phase. The TEM images revealed the
formation of larger (100–200 nm) particles with good crystallinity. Bandgap was determined to be
1.28 eV.
Most recently, the synthesis of Cu2FeSnS4 particles having a flower-like morphology has been
carried out by using a microwave irradiation method (Guan et al. 2014b). Characterization of as-
prepared CFTS nanoparticles was carried out by p-XRD, SEM, TEM, and UV-vis-NIR spectros-
copy. The CFTS nanoparticles possessed a stannite structure. The SEM images show flower-like
morphology, whereas TEM images revealed that the individual particles had a spherical structure.
The bandgap of the CFTS particles was found to be 1.52 eV, indicating its potential for solar cell
applications.
Shibuya et al. (2014) have carried out first principles calculations to investigate the substitution
of Zn atoms by Fe atoms into the CZTS lattice to form the Cu2(Zn,Fe)SnS4 solid solution in order
to tune the lattice parameters and bandgap energy. It was demonstrated that by the incorporation
of Fe atoms, a phase transition from kesterite (Zn-rich) to stannite (Fe-rich) occurred at the Fe/
Zn ratio of 0:4. An increase in bandgap and slight decrease in lattice volume were observed with
increasing Fe concentrations in the solid solution. A linear bandgap bowing was observed for each
phase, which resulted in blue-shifted photoabsorption for alloys having Fe-rich composition due
to the confinement of the conduction states. Similarly, Khadka and Kim (2014) have reported the
fabrication of polycrystalline Cu2ZnzFe1–zSnS4 (CZFTS) thin films by the chemical spray pyrolysis
process followed by sulfurization. It was found that the postsulfurized CZFTS films possess mor-
phology, structure, and optoelectronic features that make them potential photoabsorbers in thin-film
photovoltaics. Raman spectroscopy and p-XRD studies clearly showed transition from the stannite
to the kesterite phase with the increase of zinc content in the CZFTS alloy. A parabolic increasing
trend was observed in the bandgap of postsulfurized CZFTS films, and it was possible to tune the
bandgap from 1.36 ± 0.02 to 1.51 ± 0.02 eV with increasing Zn content (0 ≤ z ≤ 1).
The synthesis of Cu2MnSnS4 nanocrystals having zinc blende and wurtzite structure has been
carried out using a hot-injection approach (Liang et al. 2012b). It was found that the reactivity of the
sulfur precursor used in the reaction controls the crystal structure of Cu2MnSnS4 nanocrystals. A more
stable wurtzite nanostructure was obtained by using less reactive thiourea as the sulfur source. The
TEM images revealed that the average size of triangular zinc blende Cu2MnSnS4 nanocrystals was
9.3 nm, whereas wurtzite Cu2MnSnS4 was composed of hexagonal nanoplates having diameter of about
23.1  nm. The EDX measurements indicated that the nanocrystals had a stoichiometric composition
closely matching the theoretical composition of Cu2MnSnS4. The bandgap of the nanocrystals was
found to be 1.1 eV, indicating their potential for use in the manufacture of low-cost thin-film solar cells.
The Cu2CdSnS4 nanorods were synthesized by the solvothermal method (Cui et al. 2012b). Then,
p-XRD studies indicated a wurtzite crystallographic structure of the nanorods. The TEM images
showed that the Cu2CdSnS4 nanorods have an average diameter and length of about 5.7 and 26 nm,
respectively. The EDX measurements confirmed a relative atomic ratio of 2:1:1:4 for Cu, Cd, Sn, and
S atoms. The bandgap of the as-synthesized nanorods was determined to be 1.4 eV. Photoelectrical
measurements of the spin-casted films demonstrated that the Earth-abundant Cu2CdSnS4 nanorods
have remarkable potential as a low-cost material for thin-film solar cells. Solvothermal synthesis of
Cu2CdSnS4 nanocrystals has also been reported by other research groups (Cao et al. 2012).
This chapter highlights some of the representative semiconductor nanomaterials for solar
photovoltaics. Nanomaterials find applications in solar cells owing to their unique optoelectronic
properties, large surface area, and high optical absorption coefficient and exceptional mechanisms like
quantum confinement effects. Ternary and quaternary metal chalcogenide nanocrystals of I–III–VI2
Nanomaterials for Solar Energy 245

and I2–II–IV–VI4 family of semiconductors have demonstrated good power conversion efficiencies
(PCEs), thus proving their potential for commercialization. Current research efforts are focused on
semiconductors based on Earth-abundant and nontoxic elements. However, this field still faces chal-
lenges due to contamination from binary materials and the large variety of compositional phases. It
has been established that defect states of the nanocrystals strongly influence the performance of final
devices; therefore, theoretical modeling of the defects in semiconductor systems should be consid-
ered. Recent studies involving doping of the semiconductors have shown significant improvement
in their absorption/emission and electronic properties. Similarly, the use of small molecular ligands
or simple ions for surface passivation can have an insulating effect. Future development efforts will
emphasize greener and lower temperature-processing strategies to obtain higher yield nanomaterials
offering good PCE. The challenges associated with developing size, morphology, phase, and stoichi-
ometry control during synthesis, and understanding the fundamental characteristics as well as the
industrial-level scaled-up manufacturing issues will be optimally addressed during the next decade.

REFERENCES
Abdelhady, A. L., M. A. Malik, and P. O’Brien. 2012. High-throughput route to Cu2−xS nanoparticles from
single molecular precursor. Mater. Sci. Semicond. Proc., 15: 218–221.
Abermann, S. 2013. Non-vacuum processed next generation thin film photovoltaics: Towards marketable effi-
ciency and production of CZTS based solar cells. Solar Energy, 94: 37–70.
Ahmadi, M., S. S. Pramana, L. Xi, C. Boothroyd, Y. M. Lam, and S. Mhaisalkar. 2012. Evolution pathway of
CIGSe nanocrystals for solar cell applications. J. Phys. Chem. C., 116: 8202–8209.
Ahn, S., C. Kim, J. H. Yun, J. Gwak, S. Jeong, B.-H. Ryu, et al. 2010. CuInSe2 (CIS) thin film solar cells by
direct coating and selenization of solution precursors. J. Phys. Chem. C., 114: 8108–8113.
Ahn, S., K. Kim, A. Cho, J. Gwak, J. H. Yun, K. Shin et al. 2012. CuInSe2 (CIS) thin films prepared from
amorphous Cu–In–Se nanoparticle precursors for solar cell application. ACS Appl. Mater. Interfaces,
4: 1530–1536.
Ai, L., and J. Jiang. 2012. Hierarchical porous quaternary Cu-Fe-Sn-S hollow chain microspheres: Rapid
microwave nonaqueous synthesis, growth mechanism, and their efficient removal of organic dye pollut-
ant in water. J. Mater Chem., 22: 20586–20592.
Akhavan, V. A., B. W. Goodfellow, M. G. Panthani, C. Steinhagen, T. B. Harvey, C. J. Stolle, et al. 2012.
Colloidal CIGS and CZTS nanocrystals: A precursor route to printed photovoltaics. J. Solid State
Chem., 189: 2–12.
Aldakov, D., A. Lefrancois, and P. Reiss. 2013. Ternary and quaternary metal chalcogenide nanocrystals:
Synthesis, properties and applications. J. Mater. Chem. C., 1: 3756–3776.
An, H. S., Y. Cho, S. J. Park, H. S. Jeon, Y. J. Hwang, D.-W. Kim, et al. 2014. Cocktails of paste coatings for
performance enhancement of CuInGaS2 thin-film solar cells. ACS Appl. Mater. Interfaces, 6: 888–893.
Arani, M. S., and M. S. Niasari. 2014. A facile and reliable route to prepare flower shaped lead sulfide nano-
structures from a new sulfur source. J. Indust. Eng. Chem., 20: 3141–3149.
Arul, N. S., D. Y. Yun, D. U. Lee, and T. W. Kim. 2013. Strong quantum confinement effects in kesterite
Cu2ZnSnS4 nanospheres for organic optoelectronic cells. Nanoscale, 5: 11940–11943.
Azimi, H., Y. Hou, and C. J. Brabec, 2014. Towards low-cost, environmentally friendly printed chalcopyrite
and kesterite solar cells. Energy Environ. Sci., 7: 1829–1849.
Azizian-Kalandaragh, Y., A. Khodayari, Z. Zeng, C. Garoufalis, S. Baskoutas, and L. Gontard. 2013. Strong
quantum confinement effects in SnS nanocrystals produced by ultrasound-assisted method. J. Nanopart.
Res., 15: 1388.
Barkhouse, D. A. R., O. Gunawan, T. Gokmen, T. K. Todorov, and D. B. Mitzi. 2012. Device characteristics of
a 10.1% hydrazine-processed Cu2ZnSn(Se,S)4 solar cell. Prog. Photovolt.: Res. Appl., 20: 6–11.
Bauer, E., K. L. Man, A. Pavlovska, A. Locatelli, T. O. Mentes, M. A. Nino et al. 2014. Fe3S4 (Greigite) forma-
tion by vapor-solid reaction. J. Mater. Chem. A., 2: 1903–1913.
Beal, J. H. L., P. G. Etchegoin, and R. D. Tilley. 2012. Synthesis and characterisation of magnetic iron sulfide
nanocrystals. J. Solid State Chem., 189: 57–62.
Beal, J. H. L., S. Prabakar, N. Gaston, G. B. Teh, P. G. Etchegoin, G. Williams, et al. 2011. Synthesis and
comparison of the magnetic properties of iron sulfide spinel and iron oxide spinel nanocrystals.
Chem. Mater., 23: 2514–2517.
246 Solar Energy Conversion and Storage

Bellemare, B. 2006. Solar boiling over. UtiliPoint International Article, August 9.


Benchikri, M., O. Zaberca, R. El Ouatib, B. Durand, F. Oftinger, A. Balocchi, et al. 2012. A high temperature
route to the formation of highly pure quaternary chalcogenide particles. Mater. Lett., 68: 340–343.
Bera, P., and S. I. Seok. 2012. Nanocrystalline copper sulfide of varying morphologies and stoichiometries
in a low temperature solvothermal process using a new single-source molecular precursor. Solid State
Sci., 14: 1126–1132.
Bernardini, G. P., P. Bonazzi, M. Corazza, F. Corsini, G. Mazzetti, L. Poggi, et al. 1990. New data on the
Cu2FeSnS4-Cu2ZnSnS4 pseudobinary system at 750 degree and 550 degree C. Eur. J. Mineral, 2:
219–225.
Bernardini, G. P., D. Borrini, A. Caneschi, F. Di Benedetto, D. Gatteschi, S. Ristori, et al. 2000. EPR and
SQUID magnetometry study of Cu2FeSnS4 (stannite) and Cu2ZnSnS4 (kesterite). Phys. Chem. Miner.,
27: 453–461.
Biacchi, A. J., D. D. Vaughn, and R. E. Schaak. 2013. Synthesis and crystallographic analysis of shape-
controlled SnS nanocrystal photocatalysts: Evidence for a pseudotetragonal structural modification.
J. Am. Chem. Soc., 135: 11634–11644.
Brockway, L. O. 1934. The crystal structure of stannite, Cu2FeSnS4. Z. Kristallogr., 89: 434–441.
Brown, G. F., and J. Wu. 2009. Third generation photovoltaics. Laser Photonics Rev., 3: 394–405.
Cabán-Acevedo, M., M. S. Faber, Y. Tan, R. J. Hamers, and S. Jin. 2012. Synthesis and properties of semicon-
ducting iron pyrite (FeS2) nanowires. Nano Lett., 12: 1977–1982.
Cao, M., L. Li, W. Z. Fan, X. Y. Liu, Y. Sun, and Y. Shen. 2012. Quaternary Cu2CdSnS4 nanoparticles synthe-
sized by a simple solvothermal method. Chem. Phys. Lett., 534: 34–37.
Cao, Y., Y. Xiao, J.-Y. Jung, H.-D. Um, S.-W. Jee, H. M. Choi, et al. 2013. Highly electrocatalytic
Cu2ZnSn(S1–xSex)4 counter electrodes for quantum-dot-sensitized solar cells. ACS Appl. Mater.
Interfaces, 5: 479–484.
Cattley, C. A., C. Cheng, F. S. M. Airclough, L. M. Droessler, N. P. Young, J. H. Warner, et al. 2013. Low
temperature phase selective synthesis of Cu2ZnSnS4 quantum dots. Chem. Commun., 49: 3745–3747.
Chang, J., and E. R. Waclawik. 2013. Controlled synthesis of CuInS2, Cu2SnS3 and Cu2ZnSnS4 nano-structures:
Insight into the universal phase-selectivity mechanism. Cryst. Eng. Comm., 15: 5612–5619.
Chang, S.-H., M.-Y. Chiang, C.-C. Chiang, F.-W. Yuan, C.-Y. Chen, B.-C. Chiu, et al. 2011. Facile colloidal
synthesis of quaternary CuIn1–xGa x(SySe1–y)2 (CIGSSe) nanocrystal inks with tunable band gaps for use
in low-cost photovoltaics. Energy Environ. Sci., 4: 4929–4932.
Chapin, D. M., C. S. Fuller, and G. L. Pearson. 1954. A new silicon p-n junction photocell for converting solar
radiation into electrical power. J. Appl. Phys., 25: 676–677.
Chen, B., H. Zhong, M. Wang, R. Liu, and B. Zou. 2013b. Integration of CuInS2-based nanocrystals for high
efficiency and high colour rendering white light-emitting diodes. Nanoscale, 5: 3514–3519.
Chen, Z., M. Tang, L. Song, G. Tang, B. Zhang, L. Zhang, et al. 2013a. In situ growth of CuInS2 nanocrystals
on nanoporous TiO2 film for constructing inorganic/organic heterojunction solar cells. Nanoscale Res.
Lett., 8: 354.
Chernomordik, B. D., A. E. Béland, D. D. Deng, F. L. F. Rancis, and E. S. Aydil. 2014a. Microstructure evolu-
tion and crystal growth in Cu2ZnSnS4 thin films formed by annealing colloidal nanocrystal coatings.
Chem. Mater., 26: 3191–3201.
Chernomordik, B. D., A. E. Béland, N. D. Trejo, A. A. Gunawan, D. D. Deng, K. A. Mkhoyan, et al. 2014b.
Rapid facile synthesis of Cu2ZnSnS4 nanocrystals. J. Mater. Chem. A, 2: 10389–10395.
Chesman, A. S. R., N. W. Duffy, S. Peacock, L. Waddington, N. A. S. Webster, and J. J. Jasieniak. 2013a.
Non-injection synthesis of Cu2ZnSnS4 nanocrystals using a binary precursor and ligand approach. RSC
Adv., 3: 1017–1020.
Chesman, A. S. R., J. Van Embden, N. W. Duffy, N. A. S. Webster, and J. J. Jasieniak. 2013b. In situ formation
of reactive sulfide precursors in the one-pot, multigram synthesis of Cu2ZnSnS4 nanocrystals. Cryst.
Growth Design, 13: 1712–1720.
Chiang, M.-Y., S.-H. Chang, C.-Y. Chen, F.-W. Yuan, and H.-Y. Tuan, 2011. Quaternary CuIn(S1−xSex)2 nano-
crystals: Facile heating-up synthesis, band gap tuning, and gram-scale production. J. Phys. Chem. C,
115: 1592–1599.
Cho, A., H. Song, J. Gwak, Y.-J. Eo, J. H. Yun, K. Yoon, et al. 2014. A chelating effect in hybrid inks for
non-vacuum-processed CuInSe2 thin films. J. Mater. Chem. A, 2: 5087–5094.
Connor, S. T., C.-M. Hsu, B. D. Weil, S. Aloni, and Y. Cui. 2009. Phase transformation of biphasic Cu2S−
CuInS2 to monophasic CuInS2 nanorods. J. Am. Chem. Soc., 131: 4962–4966.
Coughlan, C., A. Singh, and K. M. Ryan. 2013. Systematic study into the synthesis and shape development in
colloidal CuIn xGa1–xS2 nanocrystals. Chem. Mater., 25: 653–661.
Nanomaterials for Solar Energy 247

Cui, Y., G. Wang, and D. Pan. 2012b. Synthesis and photoresponse of novel Cu2CdSnS4 semiconductor
nanorods. J. Mater. Chem., 22: 12471–12473.
Cui, Y., R. Deng, G. Wang, and D. Pan. 2012a. A general strategy for synthesis of quaternary semiconductor
Cu2MSnS4 (M = Co2+, Fe2+, Ni2+, Mn2+) nanocrystals. J. Mater. Chem., 22: 23136–23140.
Cummins, D. R., H. B. Russell, J. B. Jasinski, M. Menon, and M. K. Sunkara. 2013. Iron sulfide (FeS) nano-
tubes using sulfurization of hematite nanowires. Nano Lett., 13: 2423–2430.
Deng, M., S. Shen, Y. Zhang, H. Xu, and Q. Wang. 2014. A generalized strategy for controlled synthesis of
ternary metal sulfide nanocrystals. New J. Chem., 38: 77–83.
Dhere, N. G. 2007. Toward GW/year of CIGS production within the next decade. Sol. Energy Mater. Solar
Cells, 91: 1376–1382.
Dierick, R., B. Capon, H. Damm, S. Flamee, P. Arickx, E. Bruneel, et al. 2014. Annealing of sulfide stabilized
colloidal semiconductor nanocrystals. J. Mater. Chem. C, 2: 178–183.
Du, Y.-F., J.-Q. Fan, W.-H. Zhou, Z.-J. Zhou, J. Jiao, and S.-X. Wu. 2012b. One-step synthesis of stoichio-
metric Cu2ZnSnSe4 as counter electrode for dye-sensitized solar cells. ACS Appl. Mater. Interfaces, 4:
1796–1802.
Du, Y., Z. Yin, J. Zhu, X. Huang, X.-J. Wu, Z. Zeng, et al. 2012a. A general method for the large-scale synthe-
sis of uniform ultrathin metal sulphide nanocrystals. Nat. Commun., 3: 1177.
Fan, F.-J., L. Wu, and S.-H. Yu, 2014. Energetic I-III-VI2 and I2-II-IV-VI4 nanocrystals: Synthesis, photovoltaic
and thermoelectric applications. Energy Environ. Sci., 7: 190–208.
Flynn, B., I. Braly, P. A. Glover, R. P. Oleksak, C. Durgan, and G. S. Herman, 2013. Continuous flow mesoflu-
idic synthesis of Cu2ZnSnS4 nanoparticle inks. Mater. Lett., 107: 214–217.
Flynn, B., W. Wang, C.-H. Chang, and G. S. Herman. 2012. Microwave assisted synthesis of Cu2ZnSnS4 col-
loidal nanoparticle inks. Physica Status Solidi (A), 209: 2186–2194.
Ford, G. M., Q. Guo, R. Agrawal, and H. W. Hillhouse. 2011. Earth abundant element Cu2Zn(Sn1−xGex)
S4 nanocrystals for tunable band gap solar cells: 6.8% efficient device fabrication. Chem. Mater., 23:
2626–2629.
Fthenakis, V. 2009. Sustainability of photovoltaics: The case for thin-film solar cells. Renew. Sust. Energy
Rev., 13: 2746–2750.
Guan, H., Y. Shi, B. Jiao, X. Wang, and F. Yu. 2014b. Flower-like Cu2FeSnS4 particles synthesized by micro-
wave irradiation method. Chalcogenide Lett., 11: 9–12.
Guan, Z., W. Luo, and Z. Zou. 2014a. Formation mechanism of ZnS impurities and their effect on photoelec-
trochemical properties on a Cu2ZnSnS4 photocathode. Cryst. Eng. Comm., 16: 2929–2936.
Gui, Z., R. Fan, X. Chen, Y. Hu, and Z. Wang. 2004. A new colloidal precursor cooperative conversion route
to nanocrystalline quaternary copper sulfide. Mater. Res. Bull., 39: 237–241.
Guo, J., X. Wang, W.-H. Zhou, Z.-X. Chang, X. Wang, Z.-J. Zhou, et al. 2013. Efficiency enhancement of dye-
sensitized solar cells (DSSCs) using ligand exchanged CuInS2 NCs as counter electrode materials. RSC
Adv., 3: 14731–14736.
Guo, Q., G. M. Ford, W.-C. Yang, C. J. Hages, H. W. Hillhouse, and R. Agrawal. 2012. Enhancing the perfor-
mance of CZTSSe solar cells with Ge alloying. Sol. Energy Mater. Solar Cells, 105: 132–136.
Guo, Q., G. M. Ford, W.-C. Yang, B. C. Walker, S. E. A. Tach, H. W. Hillhouse, et al. 2010. Fabrication of 7.2%
efficient CZTSSe solar cells using CZTS nanocrystals. J. Am. Chem. Soc., 132: 17384–17386.
Guo, Q., H. W. Hillhouse, and R. Agrawal. 2009. Synthesis of Cu2ZnSnS4 nanocrystal ink and its use for solar
cells. J. Am. Chem. Soc. 131: 11672–11673.
Gusain, M., P. Kumar, and R. Nagarajan. 2013. Wurtzite CuInS2: Solution based one pot direct synthesis and
its doping studies with non-magnetic Ga3+ and magnetic Fe3+ ions. RSC Adv., 3: 18863–18871.
Han, S.-K., M. Gong, H.-B. Yao, Z.-M. Wang, and S.-H. Yu. 2012. One-pot controlled synthesis of hexagonal-
prismatic Cu1.94S-ZnS, Cu1.94S-ZnS-Cu1.94S, and Cu1.94S-ZnS-Cu1.94S-ZnS-Cu1.94S heteronanostructures.
Angew. Chem. Int. Ed., 51: 6365–6368.
Han, Z., D. Zhang, Q. Chen, R. Hong, C. Tao, Y. Huang, et al. 2014. Synthesis of single phase chalcopyrite
CuIn1−xGa xSe2 (0 ≤ x ≤ 1) nanoparticles by one-pot method. Mater. Res. Bull., 51: 302–308.
Harvey, T. B., I. Mori, C. J. Stolle, T. D. Bogart, D. P. Ostrowski, M. S. Glaz, et al. 2013. Copper indium gal-
lium selenide (CIGS) photovoltaic devices made using multistep selenization of nanocrystal films. ACS
Appl. Mater. Interfaces, 5: 9134–9140.
Hsin, C.-L., W.-F. Lee, C.-T. Huang, C.-W. Huang, W.-W. Wu, and L.-J. Chen. 2011. Growth of CuInSe2 and
In2Se3/CuInSe2 nano-heterostructures through solid state reactions. Nano Lett., 11: 4348–4351.
Hsu, C.-H., C.-H. Chen, and D.-H. Chen. 2013a. Decoration of PbS nanoparticles on Al-doped ZnO nanorod
array thin film with hydrogen treatment as a photoelectrode for solar water splitting. J. Alloys Compds.,
554: 45–50.
248 Solar Energy Conversion and Storage

Hsu, W.-H., H.-I. Hsiang, C.-T. Chia, and F.-S. Yen. 2013b. Controlling morphology and crystallite size of
Cu(In0.7Ga0.3)Se2 nano-crystals synthesized using a heating-up method. J. Solid State Chem., 208: 1–8.
Hu, H., C. Deng, K. Zhang, M. Sun, H. Xuan, and S. Kong. 2013a. Novel two-dimensional lead sulfide quad-
rangular pyramid-aggregated arrays with self-supporting structure prepared at room temperature.
Mater. Sci. Semicond. Proc., 16: 1566–1572.
Hu, X., G. Song, W. Li, Y. Peng, L. Jiang, Y. Xue, et al. 2013b. Phase-controlled synthesis and photocatalytic
properties of SnS, SnS2 and SnS/SnS2 heterostructure nanocrystals. Mater. Res. Bull., 48: 2325–2332.
Huang, C., Y. Chan, F. Liu, D. Tang, J. Yang, Y. Lai, et al. 2013. Synthesis and characterization of multi-
component Cu2(FexZn1–x)SnS4 nanocrystals with tunable band gap and structure. J. Mater. Chem. A, 1:
5402–5407.
Intergovernmental Panel on Climate Change. 2007. IPCC Fourth Assessment Report (AR4). IPCC.
Iqbal, M. Z., F. Wang, M. Y. Rafique, S. Ali, M. H. Farooq, and M. Ellahi. 2013. Hydrothermal synthesis,
characterization and hydrogen storage of SnS nanorods. Mater. Lett., 106: 33–36.
Jackson, P., D. Hariskos, E. Lotter, S. Paetel, R. Wuerz, R. Menner, et al. 2011. New world record efficiency for
Cu(In,Ga)Se2 thin-film solar cells beyond 20%. Prog. Photovolt.: Res. Appl., 19: 894–897.
Jiang, C., J.-S. Lee, and D. V. Talapin. 2012c. Soluble precursors for CuInSe2, CuIn1–xGa xSe2, and
Cu2ZnSn(S,Se)4 based on colloidal nanocrystals and molecular metal chalcogenide surface ligands.
J. Am. Chem. Soc., 134: 5010–5013.
Jiang, D., W. Hu, H. Wang, B. Shen, and Y. Deng. 2012a. Controlled synthesis of hierarchical CuS architec-
tures by a recrystallization growth process in a microemulsion system. J. Mater. Sci., 47: 4972–4980.
Jiang, D., W. Hu, H. Wang, B. Shen, and Y. Deng. 2012b. Synthesis, formation mechanism and photocata-
lytic property of nanoplate-based copper sulfide hierarchical hollow spheres. Chem. Eng. J., 189–190:
443–450.
Jiang, H., P. Dai, Z. Feng, W. Fan, and J. Zhan. 2012d. Phase selective synthesis of metastable orthorhombic
Cu2ZnSnS4. J. Mater. Chem., 22: 7502–7506.
Jiang, X., W. Xu, R. Tan, W. Song, and J. Chen. 2013. Solvothermal synthesis of highly crystallized quaternary
chalcogenide Cu2FeSnS4 particles. Mater. Lett., 102–103: 39–42.
Jin, C., P. Ramasamy, and J. Kim. 2014. Facile hot-injection synthesis of stoichiometric Cu2ZnSnSe4 nano-
crystals using bis(triethylsilyl) selenide. Dalton Trans., 43: 9481–9485.
Kameyama, T., T. Osaki, K.-I. Okazaki, T. Shibayama, A. Kudo, S. Kuwabata, et al. 2010. Preparation and
photoelectrochemical properties of densely immobilized Cu2ZnSnS4 nanoparticle films. J. Mater.
Chem., 20: 5319–5324.
Kawawaki, T., and T. Tatsuma. 2013. Enhancement of PbS quantum dot-sensitized photocurrents using plas-
monic gold nanoparticles. Phys. Chem. Chem. Phys., 15: 20247–20251.
Kazmerski, L. L. 2006. Solar photovoltaics R & D at the tipping point: A 2005 technology overview.
J. Electron Spectros. Relat. Phenomena 150(2–3): 105–135.
Khadka, D. B., and J. Kim. 2014. Structural transition and band gap tuning of Cu2(Zn,Fe)SnS4 chalcogenide
for photovoltaic application. J. Phys. Chem. C, 118: 14227–14237.
Khan, A. H., U. Thupakula, A. Dalui, S. Maji, A. Debangshi, and S. Acharya. 2013a. Evolution of long
range bandgap tunable lead sulfide nanocrystals with photovoltaic properties. J. Phys. Chem. C, 117:
7934–7939.
Khan, M. A. M., S. Kumar, and M. S. Al-Salhi. 2013b. Synthesis and characteristics of spray deposited CuInS2
nanocrystals thin films for photovoltaic applications. Mater. Res. Bull., 48: 4277–4282.
Khare, A., A. W. Wills, L. M. Ammerman, D. J. Norris, and E. S. Aydil. 2011. Size control and quantum con-
finement in Cu2ZnSnS4 nanocrystals. Chem. Commun., 47: 11721–11723.
Kim, K.-J., R. P. Oleksak, C. Pan, M. W. Knapp, P. B. Kreider, G. S. Herman, et al. 2014. Continuous
synthesis of colloidal chalcopyrite copper indium diselenide nanocrystal inks. RSC Adv., 4:
16418–16424.
Kim, Y., K. Woo, I. Kim, Y. S. Cho, S. Jeong, and J. Moon. 2013. Highly concentrated synthesis of copper-
zinc-tin-sulfide nanocrystals with easily decomposable capping molecules for printed photovoltaic
applications. Nanoscale, 5: 10183–10188.
Kirkeminde, A., B. A. Ruzicka, R. Wang, S. Puna, H. Zhao, and S. Ren. 2012. Synthesis and optoelectronic
properties of two-dimensional FeS2 nanoplates. ACS Appl. Mater. Interfaces, 4: 1174–1177.
Knutson, T. R., P. J. Hanson, E. S. Aydil, and R. L. Penn. 2014. Synthesis of Cu2ZnSnS4 thin films directly
onto conductive substrates via selective thermolysis using microwave energy. Chem. Commun., 50:
5902–5904.
Kreider, P. B., K.-J. Kim, and C.-H. Chang. 2014. Two-step continuous-flow synthesis of CuInSe2 nanopar-
ticles in a solar microreactor. RSC Adv., 4: 13827–13830.
Nanomaterials for Solar Energy 249

Kumarakuru, H., M. J. Coombes, J. H. Neethling, and J. E. Westraadt. 2014. Fabrication of Cu2S nanoneedles
by self-assembly of nanoparticles via simple wet chemical route. J. Alloys Compd, 589: 67–75.
Kundu, J., and D. Pradhan. 2014. Controlled synthesis and catalytic activity of copper sulfide nanostructured
assemblies with different morphologies. ACS Appl. Mater. Interfaces, 6: 1823–1834.
Lauth, J., J. Marbach, A. Meyer, S. Dogan, C. Klinke, A. Kornowski, et al. 2014. Virtually bare nanocrystal
surfaces: Significantly enhanced electrical transport in CuInSe2 and CuIn1−xGa xSe2 thin films upon
ligand exchange with thermally degradable 1-ethyl-5-thiotetrazole. Adv. Funct. Mater., 24: 1081–1088.
Li, C., E. Ha, W.-L. Wong, C. Li, K.-P. Ho, and K.-Y. Wong. 2012b. A facile arrested precipitation method for
synthesis of pure wurtzite Cu2ZnSnS4 nanocrystals using thiourea as a sulfur source. Mater. Res. Bull.,
47: 3201–3205.
Li, J., Z. Jin, T. Liu, J. Wang, D. Wang, J. Lai, et al. 2013c. Ternary and quaternary chalcopyrite Cu(In1–xGa x)Se2
nanocrystals: Organoalkali-assisted diethylene glycol solution synthesis and band-gap tuning. Cryst.
Eng. Comm., 15: 7327–7338.
Li, L., X. Liu, J. Huang, M. Cao, S. Chen, Y. Shen, et al. 2012c. Solution-based synthesis and characterization
of Cu2FeSnS4 nanocrystals. Mater. Chem. Phys., 133: 688–691.
Li, M., W.-H. Zhou, J. Guo, Y.-L. Zhou, Z.-L. Hou, J. Jiao, et al. 2012a. Synthesis of pure metastable wurtzite
CZTS nanocrystals by facile one-pot method. J. Phys. Chem. C, 116: 26507–26516.
Li, Q., C. Zou, L. Zhai, L. Zhang, Y. Yang, X. A. Chen, et al. 2013a. Synthesis of wurtzite CuInS2 nanowires
by Ag2S-catalyzed growth. Cryst. Eng. Comm., 15: 1806–1813.
Li, Q., L. Zhai, C. Zou, X. Huang, L. Zhang, Y. Yang, et al. 2013b. Wurtzite CuInS2 and CuIn xGa1–xS2 nanorib-
bons: Synthesis, optical and photoelectrical properties. Nanoscale, 5: 1638–1648.
Li, W., T. Dittrich, F. Jäckel, and J. Feldmann. 2014. Optical and electronic properties of pyrite nanocrystal
thin films: The role of ligands. Small, 10: 1194–1201.
Li, Y., Y. Hu, H. Zhang, Y. Shen, and A. Xie. 2013d. Controlled synthesis of bionic microstructures PbS
crystals by mixed cationic/anionic surfactants. Russ. J. Phys. Chem. 87A: 1239–1245.
Liang, B. Y. J., Y.-M. Shen, S.-C. Wang, and J.-L. Huang. 2013. The influence of reaction temperatures and
volume of oleic acid to synthesis SnS nanocrystals by using thermal decomposition method. Thin Solid
Films, 549: 159–164.
Liang, X. L., and X. H. Wei. 2012. Synthesis and characterization of Cu2FeSnS4 semiconductor nanocrystals
with zincblende structure. Adv. Mater. Res., 512–515: 2019–2022.
Liang, X., P. Guo, G. Wang, R. Deng, D. Pan, and X. Wei. 2012b. Dilute magnetic semiconductor Cu2MnSnS4
nanocrystals with a novel zincblende and wurtzite structure. RSC Adv., 2: 5044–5046.
Liang, X., X. Wei, and D. Pan. 2012a. Dilute magnetic semiconductor Cu2FeSnS4 nanocrystals with a novel
zincblende structure. J. Nanomater., 708648.
Liao, H.-C., M.-H. Jao, J.-J. Shyue, C. Y.-F. Hen, and W.-F. Su. 2013. Facile synthesis of wurtzite copper-zinc-
tin sulfide nanocrystals from plasmonic djurleite nuclei. J. Mater. Chem. A, 1: 337–341.
Lim, Y. S., H.-S. Kwon, J. Jeong, J. Y. Kim, H. Kim, M. J. Ko, et al. 2013. Colloidal solution-processed
CuInSe2 solar cells with significantly improved efficiency up to 9% by morphological improvement.
ACS Appl. Mater. Interfaces, 6: 259–267.
Liu, L., R. Hu, W.-C. Law, I. Roy, J. Zhu, L. Ye, et al. 2013c. Optimizing the synthesis of red- and near-infrared
CuInS2 and AgInS2 semiconductor nanocrystals for bioimaging. Analyst, 138: 6144–6153.
Liu, W., B. Guo, C. Mak, A. Li, X. Wu, and F. Zhang. 2013b. Facile synthesis of ultrafine Cu2ZnSnS4 nano-
crystals by hydrothermal method for use in solar cells. Thin Solid Films, 535: 39–43.
Liu, W. C., B. L. Guo, X. S. Wu, F. M. Zhang, C. L. Mak, and K. H. Wong. 2013a. Facile hydrothermal
synthesis of hydrotropic Cu2ZnSnS4 nanocrystal quantum dots: Band-gap engineering and phonon
confinement effect. J. Mater. Chem. A, 1: 3182–3186.
Liu, W., D. B. Mitzi, M. Yuan, A. J. Kellock, S. J. Chey, and O. Gunawan. 2009. 12% efficiency CuIn(Se,S)2
photovoltaic device prepared using a hydrazine solution process. Chem. Mater. 22: 1010–1014.
Liu, Y., D. Yao, L. Shen, H. Zhang, Z. X. Hang, and B. Yang. 2012. Alkylthiol-enabled Se powder dissolution
in oleylamine at room temperature for the phosphine-free synthesis of copper-based quaternary selenide
nanocrystals. J. Am. Chem. Soc., 134: 7207–7210.
Liu, Y., G. M.e, Y. Yue, Y. Sun, Y. Wu, X. Chen, et al. 2011. Colloidal Cu2ZnSnS4 nanocrystals generated by a
facile route using ethylxanthate molecular precursors. Phys. Status Solidi (RRL), 5: 113–115.
Liu, Y., Y. Xie, H. Cui, W. Zhao, C. Yang, Y. Wang, et al. 2013d. Preparation of monodispersed CuInS2 nano-
pompons and nanoflake films and application in dye-sensitized solar cells. Phys. Chem. Chem. Phys.,
15: 4496–4499.
Lunhong, A., and J. Jing. 2012. Self-sacrificial templating synthesis of porous quaternary Cu–Fe–Sn–S semi-
conductor nanotubes via microwave irradiation. Nanotechnology, 23: 495601.
250 Solar Energy Conversion and Storage

Macpherson, H. A., and C. R. Stoldt. 2012. Iron pyrite nanocubes: Size and shape considerations for photovol-
taic application. ACS Nano, 6: 8940–8949.
Maji, S. K., A. K. Dutta, P. Biswas, D. N. Srivastava, P. Paul, A. Mondal, et al. 2012. Synthesis and characterization
of FeS nanoparticles obtained from a dithiocarboxylate precursor complex and their photocatalytic, elec-
trocatalytic and biomimic peroxidase behavior. Appl. Catal. A: Gen., 419–420: 170–177.
Mali, S. S., H. Kim, C. S. Shim, P. S. Patil, and C. K. Hong. 2013. Polyvinylpyrrolidone (PVP) assisted single-
step synthesis of kesterite Cu2ZnSnS4 nanoparticles by solvothermal process. Phys. Status Solidi (RRL),
7: 1050–1054.
Malik, M. A., P. O’Brien, and N. Revaprasadu. 1999. A novel route for the preparation of CuSe and CuInSe2
nanoparticles. Adv. Mater., 11: 1441–1444.
Malik, S. N., S. Mahboob, N. Haider, M. A. Malik, and P. O’Brien. 2011. A colloidal synthesis of CuInSe2,
CuGaSe2 and CuIn1–xGa xSe2 nanoparticles from diisopropyldiselenophosphinatometal precursors.
Nanoscale, 3: 5132–5139.
Mariama, C. S., W. Lingling, and Z. Xintong. 2013. Easy hydrothermal preparation of Cu2ZnSnS4 (CZTS)
nanoparticles for solar cell application. Nanotechnology, 24: 495401.
McPhail, M. R., and E. A. Weiss. 2014. Role of organosulfur compounds in the growth and final surface chem-
istry of PbS quantum dots. Chem. Mater., 26: 3377–3384.
Midgett, A. G., J. M. Luther, J. T. Stewart, D. K. Smith, L. A. Padilha, V. I. Klimov, et al. 2013. Size and com-
position dependent multiple exciton generation efficiency in PbS, PbSe, and PbSxSe1–x alloyed quantum
dots. Nano Lett., 13: 3078–3085.
Mocanu, A., E. Rusen, A. Diacon, and A. Dinescu. 2014. Hierarchical nanostructures of PbS obtained in the
presence of water soluble polymers. Powder Technol., 253: 237–241.
Morrish, R., R. Silverstein, and C. A. Wolden. 2012. Synthesis of stoichiometric FeS2 through plasma-assisted
sulfurization of Fe2O3 nanorods. J. Am. Chem. Soc., 134: 17854–17857.
Murali, B., and S. B. Krupanidhi. 2013. Facile synthesis of Cu2CoSnS4 nanoparticles exhibiting red-edge-effect:
Application in hybrid photonic devices. J. Appl. Phys., 114: 144312.
Nadar, L., N. Destouches, N. Crespo-Monteiro, R. Sayah, F. Vocanson, S. Reynaud, et al. 2013. Multicolor photo-
chromism of silver-containing mesoporous films of amorphous or anatase TiO2. J. Nanopart. Res., 15: 1–10.
Nakashima, S., K. Kikushima, and K. Mukai. 2013. Infrared emitting property and spherical symmetry of
colloidal PbS quantum dots. J. Crystal Growth, 378: 537–541.
Oleksak, R. P., B. T. Flynn, D. M. Schut, and G. S. Herman. 2014. Microwave-assisted synthesis of CuInSe2
nanoparticles in low-absorbing solvents. Phys. Status Solidi. (A), 211: 219–225.
Pal, M., N. R. Mathews, R. S. Gonzalez, and X. Mathew. 2013. Synthesis of Cu2ZnSnS4 nanocrystals by sol-
vothermal method. Thin Solid Films, 535: 78–82.
Pan, D., X. Wang, Z. H. Zhou, W. Chen, C. Xu, and Y. Lu, 2009. Synthesis of quaternary semiconductor nano-
crystals with tunable band gaps. Chem. Mater., 21: 2489–2493.
Pan, J., A. A. O. El-Ballouli, L. Rollny, O. Voznyy, V. M. Burlakov, A. Goriely, et al. 2013. Automated syn-
thesis of photovoltaic-quality colloidal quantum dots using separate nucleation and growth stages. ACS
Nano, 7: 10158–10166.
Panthani, M. G., V. Akhavan, B. Goodfellow, J. P. Schmidtke, L. Dunn, A. Dodabalapur, et al. 2008. Synthesis
of CuInS2, CuInSe2, and Cu(In xGa1–x)Se2 (CIGS) nanocrystal “inks” for printable photovoltaics. J. Am.
Chem. Soc., 130: 16770–16777.
Park, J.-Y., J. H. Noh, T. N. Mandal, S. H. Im, Y. Jun, and S. I. Seok. 2013. Quaternary semiconductor
Cu2FeSnS4 nanoparticles as an alternative to Pt catalysts. RSC Adv., 3: 24918–24921.
Peng, H., G. Ma, J. Mu, K. Sun, and Z. Lei. 2014. Controllable synthesis of CuS with hierarchical structures
via a surfactant-free method for high-performance supercapacitors. Mater. Lett., 122: 25–28.
Ramasamy, K., V. L. Kuznetsov, K. Gopal, M. M. A. Alik, J. Raftery, P. P. Edwards, et al. 2013b. Organotin
dithiocarbamates: Single-source precursors for tin sulfide thin films by aerosol-assisted chemical vapor
deposition (AACVD). Chem. Mater., 25: 266–276.
Ramasamy, K., M. A. Malik, and P. O’Brien. 2012. Routes to copper zinc tin sulfide Cu2ZnSnS4 a potential
material for solar cells. Chem. Commun., 48: 5703–5714.
Ramasamy, K., M. A. Malik, N. Revaprasadu, and P. O’Brien, 2013a. Routes to nanostructured inorganic
materials with potential for solar energy applications. Chem. Mater., 25: 3551–3569.
Rath, J. K., C. Prastani, D. E. Nanu, M. Nanu, R. E. I. Schropp, A. Vetushka, et al. 2014. Fabrication of SnS quan-
tum dots for solar-cell applications: Issues of capping and doping. Physica Status Solidi (B), 251: 1309–1321.
Rath, T., W. Haas, A. Pein, R. Saf, E. Maier, B. Kunert, et al. 2012. Synthesis and characterization of copper
zinc tin chalcogenide nanoparticles: Influence of reactants on the chemical composition. Sol. Energy
Mater. Solar Cells, 101: 87–94.
Nanomaterials for Solar Energy 251

Reilly, N., M. Wehrung, R. A. O’Dell, and L. Sun. 2014. Ultrasmall colloidal PbS quantum dots. Mater. Chem.
Phys., 147: 1–4.
Riha, S. C., B. A. Parkinson, and A. L. Prieto. 2009. Solution-based synthesis and characterization of
Cu2ZnSnS4 nanocrystals. J. Am. Chem. Soc., 131: 12054–12055.
Riha, S. C., B. A. Parkinson, and A. L. Prieto. 2011. Compositionally tunable Cu2ZnSn(S1–xSex)4 nanocrystals:
Probing the effect of Se-inclusion in mixed chalcogenide thin films. J. Am. Chem. Soc., 133: 15272–15275.
Roux, F., S. Amtablian, M. Anton, G. Besnard, L. Bilhaut, P. Bommersbach, et al. 2013. Chalcopyrite thin-
film solar cells by industry-compatible ink-based process. Sol. Energy Mater. Solar Cells, 115: 86–92.
Saadeldin, M., H. S. Soliman, H. A. M. Ali, and K. Sawaby. 2014. Optical and electrical characterizations of
nanoparticle Cu2S thin films. Chin. Phys. B, 23: 046803.
Seelaboyina, R., M. Kumar, A. Venkata Madiraju, K. Taneja, A. K. Keshri, S. Mahajan, et al. 2013. Microwave
synthesis of copper indium gallium (di)selenide nanopowders for thin film solar applications. J. Renew.
Sust. Energy, 5: 031608.
Seo, Y.-H., Y. Jo, Y. Choi, K. Yoon, B.-H. Ryu, S. Ahn, et al. 2014. Thermally-derived liquid phase involv-
ing multiphase Cu(In,Ga)Se2 nanoparticles for solution-processed inorganic photovoltaic devices. RSC
Adv., 4: 18453–18459.
Shibuya, T., Y. Goto, Y. Kamihara, M. Matoba, K. Yasuoka, L. A. Burton, et al. 2014. From kesterite to stan-
nite photovoltaics: Stability and band gaps of the Cu2(Zn,Fe)SnS4 alloy. Appl. Phys. Lett., 104: 021912.
Shin, S. W., J. H. Han, C. Y. Park, A. V. Moholkar, J. Y. Lee, and J. H. Kim. 2012. Quaternary Cu2ZnSnS4 nanocrys-
tals: Facile and low cost synthesis by microwave-assisted solution method. J. Alloys Compds., 516: 96–101.
Sims, R. E. H. 2008. Hydropower, geothermal, and ocean energy. MRS Bull., 33: 389–395.
Singh, S., S. K. Samji, and M. S. R. Rao. 2012. Synthesis and characterisation of CuInGaS2 nano-ink for pho-
tovoltaic applications. J. Exp. Nanosci., 8: 320–325.
Sobhani, A., M. Salavati-Niasari, and S. M. Hosseinpour-Mashkani. 2012. Single-source molecular precursor
for synthesis of copper sulfide nanostructures. J. Cluster Sci., 23: 1143–1151.
Steinhagen, C., T. B. Harvey, C. J. Stolle, J. Harris, and B. A. Korgel. 2012. Pyrite nanocrystal solar cells:
Promising, or fool’s gold? J. Phys. Chem. Lett., 3: 2352–2356.
Steinhagen, C., M. G. Panthani, V. Akhavan, B. Goodfellow, B. Koo, and B. A. Korgel. 2009. Synthesis of
Cu2ZnSnS4 nanocrystals for use in low-cost photovoltaics. J. Am. Chem. Soc., 131: 12554–12555.
Stolle, C. J., M. G. Panthani, T. B. Harvey, V. A. Akhavan, and B. A. Korgel, 2012. Comparison of the pho-
tovoltaic response of oleylamine and inorganic ligand-capped CuInSe2 nanocrystals. ACS Appl. Mater.
Interfaces, 4: 2757–2761.
Sun, S., D. Deng, C. Kong, X. Song, and Z. Yang. 2012. Twins in polyhedral 26-facet Cu7S4 cages: Synthesis,
characterization and their enhancing photochemical activities. Dalton Trans., 41: 3214–3222.
Tang, J., S. Hinds, S. O. Kelley, and E. H. Sargent. 2008. Synthesis of colloidal CuGaSe2, CuInSe2, and
Cu(InGa)Se2 nanoparticles. Chem. Mater., 20: 6906–6910.
Thompson, M. J., T. P. A. Ruberu, K. J. Blakeney, K. V. Torres, P. S. Dilsaver, and J. Vela. 2013. Axial compo-
sition gradients and phase segregation regulate the aspect ratio of Cu2ZnSnS4 nanorods. J. Phys. Chem.
Lett., 4: 3918–3923.
Tiong, V. T., Y. Zhang, J. Bell, and H. Wang. 2014. Phase-selective hydrothermal synthesis of Cu2ZnSnS4
nanocrystals: The effect of the sulphur precursor. Cryst. Eng. Comm., 16: 4306–4313.
Todorov, T. K., K. B. Reuter, and D. B. Mitzi. 2010. High-efficiency solar cell with earth-abundant liquid-
processed absorber. Adv. Mater., 22: E156–E159.
Tsuji, I., H. Kato, and A. Kudo. 2006. Photocatalytic hydrogen evolution on ZnS−CuInS2−AgInS2 solid solu-
tion photocatalysts with wide visible light absorption bands. Chem. Mater., 18: 1969–1975.
Uhl, A. R., M. Koller, A. S. Wallerand, C. M. Fella, L. Kranz, H. Hagendorfer, et al. 2013. Cu(In,Ga)Se2
absorbers from stacked nanoparticle precursor layers. Thin Solid Films, 535: 138–142.
U.S. Energy Information Administration, U.S. Department of Energy. 2013. Int. Energy Outlook. Washington, DC.
Vinokurov, K., J. E. Macdonald, and U. Banin. 2012. Structures and mechanisms in the growth of hybrid
Ru–Cu2S nanoparticles: From cages to nanonets. Chem. Mater., 24: 1822–1827.
Wang, G., S. Wang, Y. Cui, and D. Pan. 2012. A novel and versatile strategy to prepare metal–organic molecu-
lar precursor solutions and its application in Cu(In,Ga) (S,Se)2 solar cells. Chem. Mater., 24: 3993–3997.
Wang, X., W. Zhou, Z. Chang, Z. Zhou, and S. Wu. 2013. Solvent-free synthesis of hexagonal iron sulfide
nanoflowers. Chin. J. Chem., 31: 983–986.
Wang, Y.-H. A., X. Zhang, N. Bao, B. Lin, and A. Gupta. 2011. Synthesis of shape-controlled monodisperse wurtz-
ite CuInxGa1–xS2 semiconductor nanocrystals with tunable band gap. J. Am. Chem. Soc., 133: 11072–11075.
Wei, M., Q. Du, D. Wang, L. W. Iu, G. Jiang, and C. Zhu. 2012. Synthesis of spindle-like kesterite Cu2ZnSnS4
nanoparticles using thiorea as sulfur source. Mater. Lett., 79: 177–179.
252 Solar Energy Conversion and Storage

Winkler, M. T., W. Wang, O. Gunawan, H. J. Hovel, T. K. Todorov, and D. B. Mitzi. 2014. Optical designs that
improve the efficiency of Cu2ZnSn(S,Se)4 solar cells. Energy Environ. Sci., 7: 1029–1036.
Wooten, A. J., D. J. Werder, D. J. Williams, J. L. Casson, and J. A. Hollingsworth. 2009. Solution−liquid−solid
growth of ternary Cu−in−Se semiconductor nanowires from multiple- and single-source precursors.
J. Am. Chem. Soc., 131: 16177–16188.
Xiong, S., and H. C. Zeng. 2012. Serial ionic exchange for the synthesis of multishelled copper sulfide hollow
spheres. Angew. Chem. Int. Ed., 51: 949–952.
Xu, J., C.-S. Lee, Y.-B. Tang, X. Chen, Z.-H. Chen, W.-J. Zhang, et al. 2010. Large-scale synthesis and phase trans-
formation of CuSe, CuInSe2, and CuInSe2/CuInS2 core/shell nanowire bundles. ACS Nano., 4: 1845–1850.
Yan, C., C. Huang, J. Yang, F. Liu, J. Liu, Y. Lai, et al. 2012. Synthesis and characterizations of quaternary
Cu2FeSnS4 nanocrystals. Chem. Commun., 48: 2603–2605.
Yan, X., E. Michael, S. Komarneni, J. R. Brownson, and Z.-F. Yan. 2013a. Microwave- and conventional-
hydrothermal synthesis of CuS, SnS and ZnS: Optical properties. Ceram. Int., 39: 4757–4763.
Yan, Z., Y. Zhao, M. Zhuang, J. Liu, and A. Wei. 2013b. Solvothermal synthesis of CuInS2 powders and
CuInS2 thin films for solar cell application. J. Mater. Sci.: Mater. Electronics, 24: 5055–5060.
Yang, H., L. A. Jauregui, G. Zhang, Y. P. Chen, and Y. Wu. 2012. Nontoxic and abundant copper zinc tin sulfide
nanocrystals for potential high-temperature thermoelectric energy harvesting. Nano Lett., 12: 540–545.
Yang, J., C. Bao, J. Zhang, T. Yu, H. Huang, Y. Wei, et al. 2013. In situ grown vertically oriented CuInS2
nanosheets and their high catalytic activity as counter electrodes in dye-sensitized solar cells. Chem.
Commun., 49: 2028–2030.
Yao, R. Y., Z. J. Zhou, Z. L. Hou, X. Wang, W. H. Zhou, and S. X. Wu. 2013. Surfactant-free CuInS2 nanocrys-
tals: An alternative counter-electrode material for dye-sensitized solar cells. ACS Appl. Mater. Interfaces
5: 3143–3148.
Yarema, O., D. Bozyigit, I. Rousseau, L. Nowack, M. Yarema, W. Heiss, et al. 2013. Highly luminescent, size-
and shape-tunable copper indium selenide based colloidal nanocrystals. Chem. Mater., 25: 3753–3757.
Yidong, Z., H. Weiwei, and J. Huimin. 2013. Hydrothermal fabrication of chalcopyrite-type CuInS2 film and
their optical properties. Physica Scripta, 88: 015705.
Yoon, H., S. H. Na, J. Y. Choi, M. W. Kim, H. Kim, H. S. An, et al. 2014. Carbon- and oxygen-free Cu(InGa)
(SSe)2 solar cell with a 4.63% conversion efficiency by electrostatic spray deposition. ACS Appl. Mater.
Interfaces, 6: 8369–8377.
Yu, K., P. Ng, J. Ouyang, M. B. Zaman, A. Abulrob, T. N. Baral, et al. 2013. Low-temperature approach to
highly emissive copper indium sulfide colloidal nanocrystals and their bioimaging applications. ACS
Appl. Mater. Interfaces, 5: 2870–2880.
Zaberca, O., F. Oftinger, J. Y. Chane-Ching, L. Datas, A. Lafond, P. Puech, et al. 2012. Surfactant-free CZTS
nanoparticles as building blocks for low-cost solar cell absorbers. Nanotechnology, 23: 185402.
Zhai, X., H. Jia, Y. Zhang, Y. Lei, J. Wei, Y. Gao, et al. 2014. In situ fabrication of Cu2ZnSnS4 nanoflake thin
films on both rigid and flexible substrates. Cryst. Eng. Comm., 16: 6244–6249.
Zhang, J., S. Zhang, H. Zhang, Y. Zhang, Z. Zheng, and Y. Xiang. 2014. Activated selenium for promoted
formation of metal selenide nanocrystals in solvothermal synthesis. Mater. Lett., 122: 306–308.
Zhang, Q., E. Uchaker, S. L. Candelaria, and G. Cao. 2013a. Nanomaterials for energy conversion and storage.
Chem. Soc. Rev., 42: 3127–3171.
Zhang, X., N. Bao, B. Lin, and A. Gupta. 2013b. Colloidal synthesis of wurtzite Cu2CoSnS4 nanocrystals and
the photoresponse of spray-deposited thin films. Nanotechnology, 24: 105706.
Zhang, X., N. Bao, K. Ramasamy, Y.-H. A. Wang, Y. Wang, B. Lin, et al. 2012. Crystal phase-controlled syn-
thesis of Cu2FeSnS4 nanocrystals with a band gap of around 1.5 eV. Chem. Commun., 48: 4956–4958.
Zhong, H., Y. Zhou, M. Ye, Y. He, J. Ye, C. He, et al. 2008. Controlled synthesis and optical properties of col-
loidal ternary chalcogenide CuInS2 nanocrystals. Chem. Mater., 20: 6434–6443.
Zhou, H., W.-C. Hsu, H.-S. Duan, B. Bob, W. Yang, T.-B. Song, et al. 2013. CZTS nanocrystals: A promising
approach for next generation thin film photovoltaics. Energy Environ. Sci., 6: 2822–2838.
Zhou, Y.-L., W.-H. Zhou, Y.-F. Du, M. Li, and S.-X. Wu. 2011a. Sphere-like kesterite Cu2ZnSnS4 nanoparticles
synthesized by a facile solvothermal method. Mater. Lett., 65: 1535–1537.
Zhou, Y.-L., W.-H. Zhou, M. Li, Y.-F. Du, and S.-X. Wu. 2011b. Hierarchical Cu2ZnSnS4 particles for a low-
cost solar cell: Morphology control and growth mechanism. J. Phys. Chem. C, 115: 19632–19639.
Zhu, L., B. J. Richardson, and Q. Yu. 2014. Controlled colloidal synthesis of iron pyrite FeS2 nanorods and
quasi-cubic nanocrystal agglomerates. Nanoscale, 6: 1029–1037.
Zou, C., L. Zhang, D. Lin, Y. Yang, Q. Li, X. Xu, et al. 2011. Facile synthesis of Cu2ZnSnS4 nanocrystals.
Cryst. Eng. Comm., 13: 3310–3313.
12 Other Solar Cells
Rakshit Ameta

CONTENTS
12.1 Introduction........................................................................................................................... 253
12.2 Plasmonic Solar Cells............................................................................................................ 253
12.3 Hybrid Solar Cells................................................................................................................. 255
12.4 Biohybrid Solar Cell.............................................................................................................. 258
12.5 Perovskite Solar Cells............................................................................................................ 259
12.6 Miscellaneous Solar Cells.....................................................................................................260
References....................................................................................................................................... 262

12.1 INTRODUCTION
Solar cells can be divided into three different generations. The first-generation solar cells are
made from crystalline semiconductor wafers (200–300  μm), and 90% of the solar cell market
is based on these first-generation solar cells. Approximately 40% of the cost of a solar module
is due to thick silicon wafers. Second-generation solar cells are based on thin-film (1–2  μm)
technology. This film is deposited on low-cost substrates such as glass, plastic, or stainless steel.
The main focus of these solar cells is lowering the amount of material used. They contain a
variety of semiconductors like cadmium telluride, copper indium diselenide, and so on, as well
as amorphous and polycrystalline silicon. A major limitation of thin-film solar cells is their inef-
fective absorbance near bandgap, in particular, for the indirect bandgap semiconductor silicon.
Therefore, it is important to trap light inside the solar cell in order to increase the absorbance.
Third-generation solar cells are presently investigated with the goal to increase the efficiency
using second-generation SCs.
Most of the commercial solar cells are made from a refined, highly purified silicon crystal (sili-
con wafer). These silicon solar cells and complex production processes involve a high cost, and
therefore, have generated interest in alternative technologies.

12.2  PLASMONIC SOLAR CELLS


Plasmonic solar cells are a class of photovoltaic devices that converts light into electricity by
using plasmons. This is a type of thin-film solar cell normally 1–2 μm thick. Low-cost substrates
such as glass, plastic, or steel can be used in place of costly silicon to fabricate these cells.
The major problem for thin-film solar cells is limited absorption as compared to thicker solar
cells. Plasmonic cells can have more absorption by scattering light using metal nanoparticles
excited at their surface plasmon resonance (Catchpole and Polman 2008). This permits light to
be absorbed more directly without a relatively thick additional layer. It was found that Raman
scattering can be increased by using metal nanoparticles, which makes more photons available
to excite surface plasmons and, in turn, electrons, which travel through the thin-film solar cells,
thus generating a current.
Photocurrent enhancement was observed by Stuart and Hall (1998) using silver nanoparticles,
while Westphalen et al. (2000) reported an enhancement for silver clusters incorporated into indium
tin oxide and zinc-phthalocyanine solar cells. Enhanced efficiencies for ultrathin-film organic

253
254 Solar Energy Conversion and Storage

solar cells were also observed using 5 nm–diameter silver nanoparticles (Rand et al. 2004). Gold
nanoparticles were also used for scattering and absorption of light on doped silicon resulting in 80%
enhancements (Schaadt et al. 2005). Derkacs et al. (2006) also used gold nanoparticles on thin-film
silicon, giving 8% on conversion efficiency.
Another method of utilizing surface plasmons for harvesting solar energy is to have a thin film
of silicon and a thin layer of metal deposited on the lower surface. The light will travel through
the silicon and generate surface plasmons on the interface of the silicon and metal. This generates
electric fields inside of the silicon, because electric fields do not travel very far into metals. If the
electric field is strong enough, electrons produced a photocurrent. The thin film of metal must be of
nanometric-sized grooves, which act as waveguides for the incident light to excite more photons in
the silicon thin film (Ferry et al. 2008).
When a photon strikes the substrate of a solar cell, an electron and hole pair is generated. Once
the electrons and holes are separated, they will try to recombine. If the electrons can be collected
prior to this recombination, then they can be utilized as a current in an external circuit. In a solar
cell, it is important to keep a balance by minimizing this recombination, which required thin layers
and the absorption of more photon, which needs thicker-layer Tanabe (2009).
The basic functioning of plasmonic solar cells depends on scattering and absorption of light due
to the deposition of metal nanoparticles. Silicon does not absorb light appreciably. Therefore, more
light is to be scattered across the surface in order to increase the absorption. It has been found that
metal nanoparticles help in scattering the incident light across the surface of the silicon. Surface
plasmon resonance depends mainly on the density of free electrons in the particle. The order of
densities of electrons for different metals (aluminum and silver in ultraviolet, and gold and copper
in visible light) corresponds to the resonance. This resonant frequency can be shifted by variation
of the dielectric constant for the embedding medium.
Presently, first-generation solar cells do not exceed efficiencies of about 30%, while
third-generation cells can be expected to achieve efficiencies up to 40%–60%. However, in
second-generation cells, required material needs have been reduced through the use of thin-
film technology. The goal of third-generation solar cells is to increase the efficiency of second-
generation solar cells (thin film) using materials found abundantly on Earth.
Standridge et al. (2009) fabricated titanium dioxide–based dye-sensitized solar cells by incorpo-
rating corrosion-protected silver nanoparticles as plasmonic optical elements of the photoelectrode.
There was an enhancement of the dye extinction when plasmonic particles are present, and they
increase on decreasing TiO2 thickness.
Au NPs were produced by ultrafast laser ablation in liquids and were incorporated on the devices
in a single-walled nanotube (SWNT)/poly(3,4-ethylene-dioxythiophene):poly(4-styrenesulfonate)/
NP/poly(3-hexylthiophene):[6,6]-phenyl-C61-butyric acid methyl ester/Al configuration (Kymakis
et al. 2011). The incorporation of NPs leads to a power conversion efficiency improvement of 70%.
This increase has been attributed to the improved photocurrent and fill factor due to an enhanced
exciton generation rate of the photoactive layer caused by localized surface plasmon resonances of
the conduction electrons within the NPs.
Ding et al. (2011) reported that on addition of 22 vol% of core–shell particles to a 5 μm thick
TiO2 film, the energy conversion efficiency of DSSCs increases from 2.7% to 4.0%, in spite of
a more than 20% decrease in the amount of dyes adsorbed on the composite films. Plasmonic
structures of fluorine-doped tin oxide (FTO)/TiO2/NPs-Ag and FTO/NPs-Ag/TiO2 electrodes were
fabricated and used in DSSC (Lin et al. 2012). An enhancement of 60% in photocurrent as well as
improvement in photovoltage was observed. Xu et al. (2012) observed that Au@PVP not only adds
to chemical stability to iodide/triiodide electrolyte, but also to the adhesiveness to dye molecules.
They obtained a power conversion efficiency (PCE) enhancement of 30% from 3.3% to 4.3% with
incorporation of Au@PVP NPs.
Plasmonic core–shell nanoparticles (PCSNPs) can function as nanoantenna and will improve the
efficiency of dye-sensitized solar cells. It was observed that PCSNPs with a thinner shell mainly
Other Solar Cells 255

enhance the current, whereas particles with a thicker shell improve the voltage (Liu et al. 2013).
Synthesis of TiO2 branched nanorod arrays (TiO2 BNRs) with plasmonic Au nanoparticles attached
on the surface was reported by Su et al. (2013). These Au/TiO2 BNR composites exhibit high photo-
catalytic activity in photoelectrochemical water splitting. The Au/TiO2 BNRs achieved the highest
efficiency of ∼1.27%, which indicated elevated charge separation and transportation efficiencies.
The high PEC performance is mainly due to the plasmonic effect of Au nanoparticles, which
enhances the visible light absorption, together with the large surface area, efficient charge separa-
tion, and high carrier mobility of the TiO2 BNRs.
Zhang et al. (2014) fabricated silver nanoprisms (AgNPs) on indium tin oxide (ITO) by immers-
ing it in AgNPs solution for a series of immersion times. Performance of the device was the best
when the immersion time was 30 min, corresponding to AgNPs coverage of 68%. Under this condi-
tion, the device showed short-circuit current density of 10.10 mA cm–2 (18% improved), and power
conversion efficiency of 3.88% (23% improved). Poly(3,4-ethylenedioxythiophene) (PEDOT) films
incorporating gold nanoparticles have been used for counterelectrodes in dye-sensitized solar cells
by Koussi-Daoud et al. (2014). An increase in efficiency of about 130% has been reported as com-
pared to the use of PEDOT alone.
A sandwich-structured CdS-Au-TiO2 nanorod array was used as the photoanode in a photoelectro-
chemical cell for hydrogen generation via the splitting of water (Li et al. 2014a). The gold nanoparticles
sandwiched between the TiO2 nanorod and the CdS quantum dot layer play a dual role in enhancing
the solar-to-chemical energy conversion efficiency. Jiao et al. (2015) used plasmonic Ag nanopar-
ticles with vertically aligned N-doped TiO2 nanotube arrays showing improved photoelectrochemical
performance. The N-doped anatase nanotube arrays were fabricated, and cubic silver nanoparticles
(diameter 5 nm) were deposited at room temperature on TiO2 nanotubes without organic additives.
The plasmonic Ag/N-doped TiO2 composites are favorable for the separation for photoelectron–hole
pairs and increasing electron transfer resulting in enhanced photoelectrochemical performance.
Recently, Wang et al. (2015) reported the feasibility of simultaneous production of hydrogen and
electricity as well as contaminant removal from actual urban wastewater within a dye-sensitized
photoelectrochemical cell (DSPC). The photoanode in the DSPC was a novel nanostructured plas-
monic Ag/AgCl@chiral TiO2 nanofiber (Ag and AgCl nanoparticles supported on chiral TiO2 nano-
fibers). The electrolyte was actual wastewater, with added 17-β-ethinylestradiol and copper.

12.3  HYBRID SOLAR CELLS


Hybrid solar cells combine advantages of both organic and inorganic semiconductors. Hybrid pho-
tovoltaics have organic materials that consist of conjugated polymers that absorb light as the donor
and transport holes (Milliron et al. 2005). Inorganic materials in hybrid cells are used as the accep-
tor and electron transporters in the structure. The hybrid photovoltaic devices have a potential not
only for low cost but also for scalable solar power conversion.
An organic material is mixed with a high electron transport material to form the photoactive
layer in hybrid solar cells (Shaheen et al. 2005). Assemblage of two materials together in a het-
erojunction-type photoactive layer will have a greater power conversion efficiency than a single
material (Saunders and Turner 2008). One of these materials acts as the photon absorber as well
as exciton donor, while the other material facilitates dissociation of exciton at the junction. Charge
is transferred and then separated after an exciton created in the donor is delocalized on a donor–
acceptor complex (Sariciftci et al. 1993).
The average distance an exciton can diffuse through a material before annihilation by recom-
bination is called exciton diffusion length. This is relatively short in polymers—that is, 5–10 nm
(Ginger and Greenham 1999). Excitons generated within this length close to an acceptor would
contribute to the photocurrent.
Increased efficiency can be achieved by increasing the interfacial surface area between the
organic and inorganic material to facilitate charge separation and by controlling the nanoscale
256 Solar Energy Conversion and Storage

lengths. The charges then remain separate and move toward the electrode before these recombine.
The three main nanoscale structures used are

• Mesoporous inorganic films infused with electron-donating organics


• Alternating inorganic–organic lamellar structures
• Nanowire structures

Mesoporous films have been used for a high-efficiency hybrid solar cell. The structure of mes-
oporous thin-film solar cells usually includes a porous inorganic material that is saturated with
an organic surfactant, which absorbs light and transfers electrons to the inorganic semiconductor.
The electrons are then transferred to the electrode. Herman et al. (2011) used alternating layers
of organic and inorganic compounds, which are controlled through electrodeposition-based self-
assembly. The lamellar structure and periodicity of the alternating organic-inorganic layers can be
controlled through solution chemistry. Larger organic surfactants that absorb more visible radiation
must be deposited between the layers of electron-accepting inorganics to obtain cells with practical
efficiencies.
Nanostructure-based solar cells using nanowires or nanotubes of inorganics surrounded by elec-
tron-donating organics utilizing self-organization processes have been developed. They offer the
advantages of directed charge transport and controlled phase separation between donor and accep-
tor materials (Weickert et al. 2011). The nanowire-based morphology offers reduced internal reflec-
tion, facile strain relaxation, and increased defect tolerance. The efficiencies of nanowire-based
solar cells have been increased with time. Now, it seems that they are one of the most promising
nanoscale solar hybrid technologies (Garnett et al. 2011).
Hybrid cell efficiency must be increased to start large-scale manufacturing. Three factors are
likely to affect efficiency:

• Bandgap should be reduced to absorb longer-wavelength (red) photons, which contain a


significant fraction of the energy in solar insolation. Current organic photovoltaics have
shown 70% quantum efficiency for shorter wavelength (blue) photons.
• Contact resistance between each layer in the device should be minimized to offer higher
fill factor and power conversion efficiency.
• Charge carrier mobility should be increased to allow the photovoltaics to have thicker
active layers while minimizing the recombination of the carrier and keeping the resistance
of the device low.

The size of nanoparticle control creates quantum confinement and allows for the tuning of optoelec-
tronic properties, such as bandgap and electron affinity. The cells also have a large surface area to
volume ratio, which presents more area for charge transfer to occur (Wu et al. 2005).
The photoactive layer can be created by mixing nanoparticles into a polymer matrix. Solar
devices based on polymer–nanoparticle composites mostly resemble polymer solar cells. Hole
mobilities are greater than electron mobilities, so the polymer phase is used to transport holes, while
nanoparticles transport electrons to the electrode. A large interfacial area between the polymer
phase and the nanoparticles is needed, which is achieved by dispersing the particles throughout the
polymer matrix. However, the nanoparticles need to be interconnected to form percolation networks
for electron transport. Efficiency is also affected by aspect ratio, geometry, and volume fraction of
the nanoparticles.
The nanoparticle bandgap should be tuned so that it matches the corresponding polymer. The
nanoparticles involved are typically colloids, and these are stabilized in solution by ligands, which
decrease efficiency because they impede interaction between the donor and nanoparticle acceptor
and decrease electron mobility. Some success has been achieved by Saunders (2012) by exchanging
the initial ligands for pyridine or another short-chain ligand.
Other Solar Cells 257

Hybrid solar cells based on dye-sensitized solar cells are fabricated by dye-absorbed inorganic
materials and organic materials. The most preferred inorganic material is TiO2 because it is a low-
cost material, is easily synthesized, and acts as an n-type semiconductor due to the donor-like oxy-
gen vacancies. Therefore, molecular sensitizers (dyes) are attached to the titania surface to collect
a greater portion of the solar spectrum. A photon is absorbed by a sensitizer molecule layer, which
induces electron injection into the conduction band of titania, thus resulting in flow of current.
A shorter diffusion length in such solar cells decreases the solar-to-energy conversion efficiency.
A variety of organic materials are attached to the titania to increase this diffusion length (or carrier
lifetime).
Some supramolecular or multifunctional sensitizers have been used to enhance carrier diffusion
length (Moser 2005). A dye chromophore has been modified by the addition of secondary electron
donors. Minority carriers (holes) diffuse to the attached electron donors to recombine. Therefore,
electron–hole recombination is retarded by the physical separation between the dye–cation moiety
and surface of TiO2. This results in an enhanced carrier diffusion length, and as a consequence, the
carrier lifetime is also increased.
A dye-sensitized mesoporous film of TiO2 can be used for making photovoltaic cells. Such a
solar cell is called a solid-state dye-sensitized solar cell. The pores in mesoporous TiO2 thin film
(2–50 μm) are filled with any solid hole-conducting material, such as p-type semiconductors or
organic hole-conducting material. Liquid electrolyte is changed with a solid charge transport
material in these cells; however, the process of electron–hole generation and recombination
remains the same. Here, electrons are injected from photoexcited dye into the conduction band
of titania, and holes are transported by a solid charge transport electrolyte to an electrode.
Attempts have been made to obtain organic materials that have a high solar-to-energy con-
version efficiency in dye-synthesized solar cells using mesoporous titania thin film (Lancelle-
Beltron et al. 2006).
Sofos et al. (2009) reported a nanostructured lamellar structure that provides an ideal design for
bulk heterojunction solar cells. This cell has ZnO and some small conducting organic molecules,
which co-assemble into alternating layers of organic and inorganic components. This highly orga-
nized structure is stabilized by π–π stacking between the organic molecules, which allows for con-
ducting pathways in both organic and inorganic layers. The thicknesses of the layers are within the
exciton diffusion length, thus minimizing recombination among charge carriers. Initial photocon-
ductivity measurements have exhibited higher values as compared with other organic, hybrid, and
amorphous silicon photoconductors. Such systems may prove to be promising and efficient hybrid
photovoltaic devices.
Bai et al. (2012) reported a method to improve the efficiency of the promising CNT-Si solar
cells to about 10% by hydrogen peroxide doping. It forms hybrid solar cells consisting of numerous
CNT-Si heterojunction cells and CNT-Si-H2O2 photoelectrochemical cells in parallel by hydrogen
peroxide doping. The quantum efficiency of the H2O2-doped solar cell was higher than that of the
original cell.
Wang and Tang (2012) reported a photoelectrochemical biofuel cell (PEBFC) generating elec-
trical energy directly from sunlight and biomass. This cell had a natural chlorophyll-sensitized
titanium dioxide film photoanode and Pt black cathode. The incident photon-to-current efficiency
(IPCE) was found to be 8.4%. Xia et al. (2011) synthesized mesoporous CdS spheres with large sur-
face areas and ordered pore-size distribution. This electrode has a conversion efficiency of 2.39%.
The good performance, low cost, and straightforward fabrication method made this mesoporous
CdS material promising for the development of effective photoelectrochemical cells. A hybrid
heterojunction and solid-state photoelectrochemical solar cell based on graphene woven fabrics
(GWFs) and silicon was fabricated by Li et al. (2014b), and significant improvement in power con-
version efficiency was achieved (11%).
Kong et al. (2014) reported a hybrid photoelectrode by using single-crystalline rutile TiO2
nanowires (NWs) inlaid with anatase TiO2 nanoparticles (NPs). This NW-NP electrode exhibited
258 Solar Energy Conversion and Storage

6.2% conversion efficiency, which corresponds to an almost 48% improvement over the efficiency
of the nanoparticle-dye sensitized solar cell (NP-DSC). This may be attributed to the synergistic
effects of the enhanced light confinement, charge collection, and dye loading. A novel photoanode
consisting of CdS/CdSe cosensitized sea urchin–shaped ZnO/TiO2-based nano/micro hybrid het-
erostructures offers better light scattering over the entire visible frequency domain as well as better
separation of photogenerated charge carriers. A remarkable photoconversion efficiency of 6.4% was
observed (Ali et al. 2014).
A hybrid photovoltaic/photoelectrochemical (PV/PEC) water-splitting device with a benchmark
solar-to-hydrogen conversion efficiency of 5.2% has been reported (Han et al. 2014). This cell con-
sists of a gradient-doped tungsten-bismuth vanadate (W:BiVO4) photoanode and a thin-film silicon
solar cell. Hybrid TiO2-nanoparticle (NP)/TiO2-SiO2 (TS) composites were prepared using rice straw
biotemplates and then were used as photoelectrodes in dye-sensitized solar cells (Jin et al. 2014). The
hybrid TiO2-NP/TS composite showed a maximum power conversion efficiency of 5.81%.

12.4  BIOHYBRID SOLAR CELL


A biohybrid solar cell is made up of a combination of organic matter (photosystem I) and inorganic
matter. These solar cells have been made by a team of researchers at Vanderbilt University. This
team used the photosystem I (a photoactive protein complex located in the thylakoid membrane) to
recreate the natural process of photosynthesis to obtain a greater efficiency in solar energy conver-
sion. These biohybrid solar cells present a new type of renewable energy (Ciesielskia et al. 2010;
Yehezkeli et al. 2012).
Multiple layers of this photosystem I gather photonic energy and convert it into chemical energy,
generating a current. This cell consists of similar nonorganic materials found in other solar cells,
with the only difference that the injected photosystem I complexes are introduced and gathered
for some days in the gold layer. The layers of photosystem I are made visible and appear as a thin
green film. This thin green film helps in improving the energy conversion, prompting attempts to
incorporate and improve different technologies. Spinach was used as the source for photosystem I.
Thylakoid membranes were isolated and purified to separate photosystem I from this membrane.
This resulted in improved electrical current compared to that of other solar cells (almost a thousand
times greater).
Biohybrid solar cells are advantageous because they convert solar energy to electricity with
almost 100% efficiency. This conversion is quite efficient as compared to 40% efficiency in tra-
ditional solar cells. Producing a biohybrid solar cell is relatively low cost because extracting the
protein from spinach and other plants is much cheaper as compared to the cost of metals required
to produce other solar cells. Although the efficiency of the biohybrid cells is much higher, there are
many disadvantages. Traditional solar cells produce more power than that currently being achieved
by biohybrid cells. Also, the life span of a biohybrid solar cell is short—a few weeks to about
9 months.
Qian et al. (2010) reported a solar-driven microbial photoelectrochemical cell (MPC) that can
produce sustainable energy through coupling the microbial catalysis of biodegradable organic mat-
ter with solar energy conversion. This MPC consists of a p-type cuprous oxide nanowire-arrayed
photocathode and an electricigen (Shewanella oneidensis MR-1)-colonizing anode. A substantial
current generation of 200 μA was achieved from the MPC device based on the synergistic effect of
the bioanode and photocathode at zero bias under illumination of 20 mW cm−2.
Current photovoltaic and photocatalytic systems are almost entirely based on semiconductor
materials. Efforts are being made to develop semiconductor/biomolecular composites for convert-
ing sunlight to electricity and fuels (Li et al. 2011). Light-induced cis/trans isomerization in the fam-
ily of merocyanine (MC) dyes offers a recyclable proton pumping ability, which can be potentially
used in hybrid bioelectronic devices (Tayebi et al. 2014). Lipid molecules play a critical role in sta-
bilizing the dye in a membrane structure for practical use in energy devices. A major modification
Other Solar Cells 259

in the cell was introduced by eliminating the I2/I− electrolyte, resulting in a twofold increase in the
open-circuit voltage as compared with that of the conventional cell. The charging time was also
reduced by approximately four orders of magnitude.

12.5  PEROVSKITE SOLAR CELLS


A perovskite solar cell is a type of solar cell that includes a perovskite absorber, most commonly
a hybrid organic–inorganic lead or tin halide–based material, as the light-harvesting active layer.
Perovskite absorber materials are extremely cheap to produce and simple to manufacture (e.g., meth-
ylammonium or formamidinium lead halide). Solar cell efficiencies of such devices have increased
from 3.8% (Kojima et al. 2009) to 20.1%. It is an advancing solar technology. Their high efficiencies
and production costs make perovskite solar cells an commercially attractive option.
The name perovskite solar cell was derived from perovskite structure. The unit cell of the
most commonly employed perovskite absorber in solar cells is methylammonium lead trihalide
(CH3NH3PbX3, where X I−, Br−, Cl−). It has a bandgap between 2.3 and 1.57 eV depending on the
content of halide. Formamidinium lead trihalide (H2NCH3NH3PbX3) is another newer material,
which shows promise, with a bandgap between 2.23 and 1.48 eV. This minimum bandgap is closer
to the optimal for a single-junction cell than methylammonium lead trihalide; therefore, it should be
capable of higher efficiencies (Eperon et al. 2014). Lead in perovskite materials may be replaced by
tin in perovskite absorber, CH3NH3SnI3, and this cell shows a power-conversion efficiency of more
than 6% (Noel et al. 2014).
Perovskite solar cells hold an advantage over traditional silicon solar cells in the simplicity of
their processing. Traditional silicon cells require expensive, multistep processes with high tem-
peratures (>1000°C) and vacuums in special clean-room facilities to produce high-purity silicon
wafers. Organic–inorganic perovskite material can be manufactured with simpler wet chemistry
and processing techniques. Most notably, methylammonium and formamidinium lead trihalides
have been prepared using a variety of solvent techniques and vapor deposition techniques. Lead
halide and methylammonium iodide can be dissolved in solvent and spin coated onto a substrate in
solution processing. Simple solution processing results in the development of voids, platelets, and
other defects in the layer, which may decrease the efficiency of a solar cell. Lee et al. (2012) men-
tioned that rapid improvement of perovskite solar cells has made them the new generation of the
photovoltaics world with high promise. There is a huge potential for more efficient perovskite solar
cells to reach in excess of 20% power conversion efficiency.
Zhang et al. (2013) reported enhanced photocurrent in perovskite solar cells utilizing plasmonic
core–shell (gold–silica) nanoparticles. Although electrical power conversion efficiencies around
17% have been obtained in organic–inorganic lead halide perovskite solar cells, the potential toxi-
cology of lead is a problem. Tin may be a suitable substitute. While organic–inorganic tin halide
perovskites have shown good semiconducting behavior, the instability of tin in its 2+ oxidation state
is an existing challenge. Completely lead free, a CH3NH3SnI3 perovskite solar cell possessing a
mesoporous TiO2 has shown efficiency over 6%.
In the last 2–3 years, a new class of solar cell, which is based on mixed organic–inorganic
halide perovskites, has emerged. The first efficient solid-state perovskite cells were reported in
2012, they attained 16.2% energy appreciable conversion efficiencies by 2013, and they were
claimed to achieve between 17.9% and 19.3% efficiencies last year. Green et al. (2014) has
reviewed this field.
Organic–inorganic perovskites, such as CH3NH3PbX3 (X = I, Br, Cl), have emerged as attrac-
tive absorber materials for the fabrication of low-cost and high-efficiency solar cells. The n-type
contact was modified with a self-assembled fullerene monolayer, and electron transfer is switched
on. Both n-type and p-type heterojunctions with the perovskite are active in driving the photovoltaic
operation. The fullerene-modified devices achieve up to 17.3% power conversion efficiency with
significantly reduced hysteresis.
260 Solar Energy Conversion and Storage

Perovskite-based solar cells using CH3NH3PbI3−xCl x with different hole-transporting materials


were fabricated (Giacomo et al. 2014). 2,2′,7,7′-Tetrakis-(N,N-di-p-methoxyphenylamine)-9,9′-
spirobifluorene (Spiro-OMeTAD) has been compared with poly(3-hexylthiophene-2,5-diyl) (P3HT).
A power conversion efficiency of 9.3% was obtained by tuning the energy level of P3HT and opti-
mizing the fabrication of the device. This is the highest reported efficiency for a solar cell using
P3HT.
Bai et al. (2014) reported high-performance planar heterojunction perovskite solar cells. PCE
up to 15.9% has been achieved in this cell. Successful fabrication of highly efficient, stable, and
reproducible planar heterojunction CH3NH3PbI3–xCl x solar cells was made using improved cath-
ode interface bilayer-structured electron-transporting interlayers of [6,6]-phenyl-C61-butyric acid-
methyl ester (PCBM)/ZnO. They also demonstrated that this simple planar structure is promising
for large-scale devices.
Moisture is assumed to be detrimental to organometal trihalide perovskite, as excess water can
damage the crystallinity of the perovskite structure. A growth mode for thermal annealing of the
perovskite precursor film in a humid environment (e.g., ambient air) has been reported (You et al.
2014) to greatly improve the film quality, grain size, carrier mobility, and lifetime. This method pro-
duces devices with maximum power conversion efficiency of 17.1% and fill factor of 80%, revealing
a promising route to achieve high-quality perovskite polycrystalline films with superior optoelec-
tronic properties that can pave the way toward efficient photovoltaic conversion.
Different lengths of rutile TiO2 nanowires with wide-open space for effective material fill-
ing were used as photoanodes by Jiang et al. (2014) for perovskite solar cells with an efficiency
of 11.7%. A uniform and pinhole free hole-blocking layer is necessary for high-performance
perovskite-based thin-film solar cells. Wu et al. (2014) investigated the effect of nanoscale pin-
holes in compact TiO2 layers on the performance of solar cells. It was shown that TiO2 compact
layers fabricated using atomic layer deposition (ALD) contain a much lower density of nanoscale
pinholes than layers obtained by spin coating and spray pyrolysis methods. This TiO2 layer acts as
an efficient hole-blocking layer in perovskite solar cells and gives a power conversion efficiency
of 12.56%.
Inorganic–organic lead halide perovskite materials appear quite promising for the next gen-
eration of solar devices due to their high power conversion efficiency. The highest efficiencies
reported for these cells so far have been obtained mainly with methylammonium lead halide
materials. A narrow bandgap, but relatively unstable formamidinium lead iodide (FAPbI3) with
methylammonium lead bromide (MAPbBr3) has been used as the light-harvesting unit in a bilayer
solar cell architecture. Incorporation of MAPbBr3 into FAPbI3 has stabilized the perovskite phase
of FAPbI3 and improved the power conversion efficiency of the solar cell to more than 18% (Jeon
et al. 2015).

12.6  MISCELLANEOUS SOLAR CELLS


A polymer solar cell is a type of flexible solar cell made with polymers. Polymer solar cells include
organic solar cells and are also called plastic solar cells. These polymer solar cells are lightweight
as compared to silicon-based devices, disposable, inexpensive to fabricate, flexible, and have rela-
tively less adverse environmental impact. These cells have the potential to exhibit transparency,
suggesting applications in windows, walls, flexible electronics, and so on, but there are some disad-
vantages of polymer solar cells. They offer about one-third of the efficiency of hard materials and
may undergo substantial photochemical degradation. Although polymer solar cells have some sta-
bility problems, these cells are low cost and more efficient. Polymer solar cells were able to achieve
over 10% efficiency via a tandem structure.
Carbon nanotubes (CNTs) have high electron and thermal conductivity, robustness, and flexibil-
ity. The CNTs have been used as either the photoinduced exciton carrier transport medium impurity
Other Solar Cells 261

within a polymer-based photovoltaic layer or as the photoactive (photon–electron conversion) layer.


Electron-accepting impurities must be added to the photoactive region to increase the photovoltaic
efficiency. Dissociation of the exciton pair can be accomplished by the CNT matrix by incorporat-
ing those into the polymer. The high surface area (~1600 m2/g) of CNTs offers a good opportunity
for exciton dissociation (Cinke et al. 2002). Separated carriers within the polymer–CNT matrix are
transported by the percolation pathways of adjacent CNTs, providing the means for high carrier
mobility and efficient charge transfer. The factors of performance of CNT–polymer hybrid photo-
voltaics are low as compared to those of inorganic photovoltaics.
Metal nanoparticles may be applied to the exterior of CNTs to increase the exciton separation
efficiency. The metal provides a higher electric field at the CNT–polymer interface, thus accelerat-
ing the exciton carriers to transfer them more effectively to the CNT matrix. The CNT may be used
as an add-in material to increase carrier transport and also as the photoactive layer (Chen et al.
2008). The single-walled CNT (SWCNT) is a potentially attractive material for photovoltaic appli-
cations because of its unique structural and electrical properties. This has high electric conductivity
and shows ballistic carrier transport, thus greatly decreasing carrier recombination. The bandgap
of the SWCNT is inversely proportional to the diameter of the tube. Therefore, SWCNT may show
multiple, direct bandgaps.
Several challenges must be addressed for CNT to be used in photovoltaic applications. CNT
degrades in an oxygen-rich environment with time. The passivation layer required to prevent CNT
oxidation may reduce the optical transparency of the electrode region; hence, the photovoltaic effi-
ciency is lower. The dispersion of CNT within the polymer photoactive layer is another problem.
It should be well dispersed within the polymer matrix to form charge-transfer-efficient pathways
between the excitons and the electrode (Somani et al. 2008).
The CNT is lacking to form a p-n junction, due to the difficulty of doping certain segments.
A p-n junction creates an internal built-in potential, which provides a pathway for efficient carrier
separation. Energy band bending using two electrodes of different work functions may solve this
problem. The oxidation of CNT will create another difficulty. Oxidized CNTs have a tendency to
become more metallic, therefore limiting their use as a photovoltaic material (Collins et al. 2000).
A quantum dot solar cell utilizes quantum dots (QDs) as the absorbing photovoltaic material.
Bulk materials such as silicon, CuInGaSe2, or CdTe are replaced by quantum dots, because their
bandgaps are tunable across a wide range of energy levels by changing the size of QDs. The band-
gap is fixed by the choice of material. This property makes quantum dots attractive for multijunc-
tion solar cells, where a variety of materials may be used to improve the efficiency by harvesting
multiple portions of the solar spectrum. In such solar cells, 8.7% efficiency has been reported by
Chuang et al. (2014).
The idea of using quantum dots for achieving high efficiency was first noted by Barnham and
Duggan (1990). Quantum dots are semiconducting particles. Their sizes have been reduced to less
than exciton Bohr radius, and as such, electron energies that can exist within them become finite,
much like energies in an atom due to quantum mechanics considerations. Quantum dots have also
been referred to as artificial atoms. They can be grown over a range of sizes, allowing them to
have a variety of bandgaps without changing the underlying material or construction techniques
(Baskoutas and Terzis 2006). This tuning of bandgap is achieved by varying the duration of syn-
thesis or temperature, and this ability makes quantum dots desirable for solar cells. Single-junction
solar cells lead sulfide (PbS) CQDs have bandgaps that can be tuned into the far-infrared, which is
very difficult to achieve with traditional materials modification. Half of the solar energy reaching
the Earth is in the infrared region, and mostly in the near-infrared region. A quantum dot solar cell
makes this infrared energy as accessible as any other (Sargent 2005).
In the future, many new types of solar cells will be developed, or modifications of existing
generations of solar cells will be made, leading to easy fabrication, higher efficiency, lower cost,
increased durability, and lighter weight.
262 Solar Energy Conversion and Storage

REFERENCES
Ali, Z., I. Shakir, and D. J. Kang. 2014. Highly efficient photoelectrochemical response by sea-urchin shaped
ZnO/TiO2 nano/micro hybrid heterostructures co-sensitized with CdS/CdSe. J. Mater. Chem. A, 2:
6474–6479.
Bai, S., Z. Wu, X. Wu, Y. Jin, N. Zhao, Z. Chen, et al. 2014. High-performance planar heterojunction perovskite
solar cells: Preserving long charge carrier diffusion lengths and interfacial engineering. Nano Res., 7:
1749–1758.
Bai, X., H. Wang, J. Wei, Y. Jia, H. Zhu, K. Wang, et al. 2012. Carbon nanotube-silicon hybrid solar cells with
hydrogen peroxide doping. Chem. Phys. Lett., 533: 70–73.
Barnham, K. W. J., and G. Duggan. 1990. A new approach to high-efficiency multi-band-gap solar cells.
J. Appl. Phys., 67: 3490.
Baskoutas, S., and A. F. Terzis. 2006. Size-dependent band gap of colloidal quantum dots. J. Appl. Phys., 99:
013708.
Catchpole, K. R., and A. Polman. 2008. Plasmonic solar cells. Opt. Express, 16: 21793–21800.
Chen, C., Y. Lu, E. S. Kong, Y. Zhang, and S.-T. Lee. 2008. Nanowelded carbon-nanotube-based solar micro-
cells. Small, 4: 1313–1318.
Chuang, C.-H. M., P. R. Brown, V. Bulović, and M. G. Bawendi. 2014. Improved performance and stability in
quantum dot solar cells through band alignment engineering. Nat. Mater., 13: 796–801.
Ciesielskia, P. N., F. M. Hijazib, A. M. Scott, C. J. Faulkner, L. Beard, K. Emmett, et al. 2010. Photosystem
I-based biohybrid photoelectrochemical cells. Biosource Technol., 101: 3047–3053.
Cinke, M., J. Li, B. Chen, A. Cassell, L. Delzeit, J. Han, et al. 2002. Pore structure of raw and purified HiPco
single-walled carbon nanotubes. Chem. Phys. Lett., 365: 69–74.
Collins, P. G., K. Bradley, M. Ishigami, and A. Zettl. 2000. Extreme oxygen sensitivity of electronic properties
of carbon nanotubes. Science, 287: 1801–1804.
Derkacs, D., S. H. Lim, P. Matheu, W. Mar, and E. T. Yu. 2006. Improved performance of amorphous silicon
solar cells via scattering from surface plasmon polaritons in nearby metallic nanoparticles. Appl. Phys.
Lett., 89: 093103.
Ding, B., B. J. Lee, M. Yang, H. S. Jung, and J.-K. Lee. 2011. Surface-plasmon assisted energy conversion in
dye-sensitized solar cells. Adv. Energy Mater., 1: 415–421.
Eperon, G. E., S. D. Stranks, C. Menelaou, M. B. Johnston, L. M. Herz, and H. J. Snaith. 2014. Formamidinium
lead trihalide: A broadly tunable perovskite for efficient planar heterojunction solar cells. Energy
Environ. Sci., 7: 982–988.
Ferry, V. E., L. A. Sweatlock, D. Pacifici, and H. A. Atwater. 2008. Plasmonic nanostructure design for effi-
cient light coupling into solar cells. Nano Lett., 8: 4391–4397.
Garnett, E. C., M. L. Brongersma, Y. Cui, and M. D. McGehee. 2011. Nanowire solar cells. Ann. Rev. Mater.
Res., 41: 269–295.
Giacomo, F. D., S. Razza, F. Matteocci, A. D’Epifanio, S. Licoccia, T. M. Brown, et al. 2014. High efficiency
CH3NH3PbI(3−x)Cl x perovskite solar cells with poly(3-hexylthiophene) hole transport layer. J. Power
Sources, 251: 152–156.
Ginger, D. S., and N. C. Greenham. 1999. Photoinduced electron transfer from conjugated polymers to CdSe
nanocrystals. Phys. Rev. B, 59: 10622–10629.
Green, M. A., A. Ho-Baillie, and H. J. Snaith. 2014. The emergence of perovskite solar cells. Nat. Photonics,
8: 506–514.
Han, L., F. F. Abdi, R. Vande Krol, R. Liu, Z. Huang, H.-J. Lewerenz, et al. 2014. Efficient water-splitting
device based on a bismuth vanadate photoanode and thin-film silicon solar cells. ChemSusChem., 7:
2832–2838.
Herman, D. J., J. E. Goldberger, S. Chao, D. T. Martin, and S. I. Stupp 2011. Orienting periodic organic−inor-
ganic nanoscale domains through one-step electrodeposition. ACS Nano., 5: 565–573.
Jeon, N. J., J. H. Noh, W. S. Yang, Y. C. Kim, S. Ryu, J. Seo, et al. 2015. Compositional engineering of
perovskite materials for high-performance solar cells. Nature, 517: 476–480.
Jiang, Q., X. Sheng, Y. Li, X. Feng, and T. Xu. 2014. Rutile TiO2 nanowire-based perovskite solar cells. Chem.
Commun., 50: 14720–14723.
Jiao, J., J. Tang, W. Gao, D. Kuang, Y. Tong, and L. Chen. 2015. Plasmonic silver nanoparticles matched with
vertically aligned nitrogen-doped titanium dioxide nanotube arrays for enhanced photoelectrochemical
activity. J. Power Sources, 274: 454–470.
Jin, E. M., J.-Y. Park, K.-J. Hwang, H.-B. Gu, and S. M. Jeong. 2014. Biotemplated hybrid TiO2 nanoparticle
and TiO2-SiO2 composites for dye-sensitized solar cells. Mater. Lett., 131: 190–193.
Other Solar Cells 263

Kojima, A., K. Teshima, Y. Shirai, and T. Miyasaka. 2009. Organometal halide perovskites as visible-light
sensitizers for photovoltaic cells. J. Am. Chem. Soc., 131: 6050–6051.
Kong, E.-H., Y.-H. Yoon, Y.-J. Chang, and H. M. Jang. 2014. Hybrid photoelectrode by using vertically aligned
rutile TiO2 nanowires inlaid with anatase TiO2 nanoparticles for dye-sensitized solar cells. Mater. Chem.
Phys., 143: 1440–1445.
Koussi-Daoud, S., D. Schaming, P. Martin, and J.-C. Lacroix. 2014. Gold nanoparticles and poly(3,4-ethylene-
dioxythiophene) (PEDOT) hybrid films as counter-electrodes for enhanced efficiency in dye-sensitized
solar cells. Electrochim. Acta, 125: 601–660.
Kymakis, E., E. Stratakis, E. Koudoumas, and C. Fotakis. 2011. Plasmonic organic photovoltaic devices on
transparent carbon nanotube films. IEEE Trans. Electron. Devices, 58: 860–864.
Lancelle-Beltran, E., P. Prené, C. Boscher, P. Belleville, P. Buvat, and C. Sanchez. 2006. All-solid-state dye-
sensitized nanoporous TiO2 hybrid solar cells with high energy-conversion efficiency. Adv. Mater., 18:
2579–2582.
Lee, M. M., J. Teuscher, T. Miyasaka, and H. J. Snaith. 2012. Efficient hybrid solar cells based on meso-
superstructured organometal halide perovskites. Science, 338: 643–647.
Li, C., F. Wang, and J. C. Yu. 2011. Semiconductor/biomolecular composites for solar energy applications.
Energy Environ. Sci., 4: 100–113.
Li, J., S. K. Cushing, P. Zheng, T. Senty, F. Meng, A. D. Bristow, et al. 2014a. Solar hydrogen generation by a
CdS-Au-TiO2 sandwich nanorod array enhanced with au nanoparticle as electron relay and plasmonic
photosensitizer, J. Am. Chem. Soc., 136: 8438–8449.
Li, X., X. Zang, X. Li, M. Zhu, Q. Chen, K. Wang, et al. 2014b. Hybrid heterojunction and solid-state photo-
electrochemical solar cells. Adv. Energy Mater., 4: 1400224.
Lin, S.-J., K.-C. Lee, J.-L. Wu, and J.-Y. Wu. 2012. Plasmon-enhanced photocurrent in dye-sensitized solar
cells. Solar Energy, 86: 2600–2605.
Liu, W.-L., F.-C. Lin, Y.-C. Yang, C.-H. Huang, S. Gwo, M. H. Huang, et al. 2013. The influence of shell thick-
ness of Au@TiO2 core-shell nanoparticles on the plasmonic enhancement effect in dye-sensitized solar
cells. Nanoscale, 5: 7953–7962.
Milliron, D. J., I. Gur, and A. P. Alivisatos. 2005. Hybrid organic–nanocrystal solar cells. MRS Bull., 30:
41–44.
Moser, J. 2005. Solar cells: Later rather than sooner. Nature Mater., 4: 723–724.
Noel, N. K., S. D. Stranks, A. Abate, C. Wehrenfennig, S. Guarnera, A.-A. Haghighirad, et al. 2014. Lead-
free organic–inorganic tin halide perovskites for photovoltaic applications. Energy Environ. Sci., 7:
3061–3068.
Qian, F., G. Wang, and Y. Li. 2010. Solar-driven microbial photoelectrochemical cells with a nanowire photo-
cathode. Nano Lett., 10: 4686–4691.
Rand, B. P., P. Peumans, and S. R. Forrest. 2004. Long-range absorption enhancement in organic tandem thin-
film solar cells containing silver nanoclusters. J. Appl. Phys., 96: 7519.
Sargent, H. E. 2005. Infrared quantum dots. Adv. Mater., 17: 515–522.
Sariciftci, N. S., L. Smilowitz, A. J. Heeger, and F. Wudl. 1993. Semiconducting polymers (as donors) and
buckminsterfullerene (as acceptor): Photoinduced electron transfer and heterojunction devices. Synth.
Met., 59: 333–352.
Saunders, B. R. 2012. Hybrid polymer/nanoparticle solar cells: Preparation, principles and challenges.
J. Colloid Interface Sci., 369: 1–15.
Saunders, B. R., and M. L. Turner. 2008. Nanoparticle-polymer photovoltaic cells. Adv. Colloid Interface Sci.,
138: 1–23.
Schaadt, D. M., B. Feng, and E. T. Yu. 2005. Enhanced semiconductor optical absorption via surface plasmon
excitation in metal nanoparticles. Appl. Phys. Lett., 86: 063106.
Shaheen, S. E., D. S. Ginley, and G. E. Jabbour. 2005. Organic-based photovoltaics. MRS Bull., 30: 10–19.
Sofos, M., J. Goldberger, D. A. Stone, J. E. Allen, Q. Ma, D. J. Herman, et al. 2009. A synergistic assembly of
nanoscale lamellar photoconductor hybrids. Nature Mater., 8: 68–75.
Somani, P. R., S. P. Somani, and M. Umeno. 2008. Application of metal nanoparticles decorated carbon nano-
tubes in photovoltaics. Appl. Phys. Lett., 93: 033315.
Standridge, S. D., G. C. Schatz, and J. T. Hupp. 2009. Distance dependence of plasmon-enhanced photocurrent
in dye-sensitized solar cells. J. Am. Chem. Soc., 131: 8407–8409.
Stuart, H. R., and D. G. Hall. 1998. Island size effects in nanoparticle-enhanced photodetectors. Appl. Phys.
Lett., 73: 3815.
Su, F., T. Wang, R. Lv, J. Zhang, P. Zhang, J. Lu, et al. 2013. Dendritic Au/TiO2 nanorod arrays for visible-light
driven photoelectrochemical water splitting. Nanoscale, 5: 9001–9009.
264 Solar Energy Conversion and Storage

Tanabe, K. 2009. A review of ultrahigh efficiency III-V semiconductor compound solar cells: Multijunction
tandem, lower dimensional, photonic up/down conversion and plasmonic nanometallic structures.
Energies, 2: 504–530.
Tayebi, L., M. Mozafari, R. El-Khouri, P. Rouhani, and D. Vashaee. 2014. Energy harvesting capability of
lipid-merocyanine macromolecules: A new design and performance model development. Photochem.
Photobiol., 90: 517–521.
Wang, D., Y. Li, G. L. Puma, C. Wang, P. Wang, W. Zhang, et al. 2015. Dye-sensitized photoelectrochemical
cell on plasmonic Ag/AgCl @ chiral TiO2 nanofibers for treatment of urban wastewater effluents, with
simultaneous production of hydrogen and electricity. Appl. Catal. B: Environ., 168–169: 25–32.
Wang, K., and J. Tang. 2012. Natural chlorophyll-sensitized nanocrystalline TiO2 as photoanode of a hybrid
photoelectrochemical biofuel cell. Adv. Mater. Res., 496: 399–402.
Weickert, J., R. B. Dunbar, W. Wiedemann, H. C. Hesse, and L. Schmidt-Mende. 2011. Nanostructured organic
and hybrid solar cells. Adv. Mater., 23: 1810–1828.
Westphalen, M., U. Kreibig, J. Rostalski, H. Lüth, and D. Meissner. 2000. Metal cluster enhanced organic
solar cells. Sol. Energy Mater. Solar Cells, 61: 97–105.
Wu, X. L., J. Y. Fan, T. Qiu, X. Yang, G. G. Siu, and P. K. Chu. 2005. Experimental evidence for the quantum
confinement effect in 3C-SiC nanocrystallites. Phys. Rev. Lett., 94: 026102.
Wu, Y., X. Yang, H. Chen, K. Zhang, C. Qin, J. Liu, et al. 2014. Highly compact TiO2 layer for efficient hole-
blocking in perovskite solar cells. Appl. Phys. Express, 7: 052301.
Xia, C., N. Wang, and X. Kim. 2011. Mesoporous CdS spheres for high-performance hybrid solar cells.
Electrochim. Acta, 56: 9504–9507.
Xu, Q., F. Liu, W. Meng, and Y. Huang. 2012. Plasmonic core-shell metal-organic nanoparticles enhanced
dye-sensitized solar cells. Opt Express, 20: A898–A907.
Yehezkeli, O., R. Tel-Vered, J. Wasserman, A. Trifonov, D. Michaeli, R. Nechushtai, et al. 2012. Integrated
photosystem II-based photoelectrochemical cells. Nature Commun., 3: 742.
You, J., Y.-(M.) Yang, Z. Hong, T.-B. Song, L. Meng, Y. Liu, et al. 2014. Moisture assisted perovskite film
growth for high performance solar cells. Appl. Phys. Lett., 105: 183902.
Zhang, Q., W. J. Qin, H. Q. Cao, L. Y. Yang, and S.-G. Yin. 2014. Efficiency enhancement of organic solar
cells with process-optimized silver nanoprisms. Appl. Mech. Mater., 598: 327–330.
Zhang, W., M. Saliba, S. D. Stranks, Y. Sun, X. Shi, U. Wiesner, et al. 2013. Enhancement of perovskite-based
solar cells employing core-shell metal nanoparticles. Nano Lett., 13: 4505–4510.
Index
A solution-processed, 73
solvent-processes, 73–74
Absorption photon-to-current efficiency (APCE), 145, 146
Air mass ratio (m), 8
3-Aminopropyltriethoxysilane (APTES), 78 C
APCE, see Absorption photon-to-current efficiency
(APCE) Carbon dioxide (CO2), 12–13
Artificial atoms, 261 in atmosphere, 173
Artificial photosynthesis, 10, 41, 155, 163, 187, 195 conversion of, to fuels, 173–174
approaches in, 190–196 emissions, 2
basic components of, 190–191 mechanism of photocatalytic reduction of, 174–176
catalysis for CO2 reduction and fuel production, molecule, characteristics of, 174
208–213 operation conditions, 181–182
catalytic proton reduction and hydrogen evolution pressure effect, 181
light-driven catalysis for hydrogen production, reactor-type effect, 182
206–207 reductant effect, 181
proton reduction catalyst by hydrogenases, temperature effect, 181
204–206 outlook, 182–183
catalytic water oxidation, 199–200 photocatalysts for reduction of, 177–181
Iridium- based catalyst, 202–203 metal oxide photocatalysts, 179
Mn-, Co-, and Fe-based catalysts, 203–204 nonoxide semiconductor photocatalysts, 179–181
Ruthenium-based catalysts, 200–202 TiO2 and related titanium-containing solids,
charge separation by molecular donor–acceptor 177–179
systems, 197–198 reduction and fuel production, 208–213
heterogeneous artificial photosynthesis systems, Carbon nanotubes (CNTs), 74–75, 100, 101, 260–261
195–196 Cascade heterojunction organic solar cells, 75
homogeneous artificial photosynthesis systems, Catalytic proton reduction and hydrogen evolution
191–195 light-driven catalysis for hydrogen production,
light absorption, 196–197 206–207
light-driven catalytic water splitting, 199 proton reduction catalyst by hydrogenases, 204–206
natural photosynthesis to, 188–190 Catalytic water oxidation, 199–204
for production of oxygen and hydrogen, 191 Iridium-based catalyst, 202–203
Au NPs, 254, 255 Mn-, Co-, and Fe-based catalysts, 203–204
Ruthenium-based catalysts, 200–202
B Cathode interfacial material (CIM), 67
Cell efficiency, 36, 99
Band bending, 21, 166 hybrid, 256
in n-doped semiconductor, 19 of photogalvanic cells, 117
in p-doped semiconductor, 19, 20 Charge
Band engineering, 182 carrier mobilities, 64, 67
Bandgaps, reducing through doping, 159 collection, 60
Becquerel, Alexandre-Edmond, 86 separation, 59, 197–198
Becquerel effect, 115 transfer, 59–60
Bell Laboratories, 1, 57, 86, 220 Charge-transfer complex, 59
Bilayer cells, 14, 69–71 Chlor-alkali plants, 153
Bilayer organic photovoltaic cells, 69–71, 72 Chlorophyll, 9, 10, 145, 189, 192, 196
Biohybrid solar cells, 14, 258–259 Cis-genes, 205
Biohydrogen, 148 CNT-Si solar cells, 257
Biological hydrogen production, 148 Composite photocatalyst, 163
Biomass waste, of wine industries, 154 Concentrated solar power (CSP), 143, 149–150
Bromophenol red-EDTA system, 124 Concentration cell, 117, 118
Bulk heterojunction (BHJ), 71–75 Conduction band (CB), 10, 11, 13, 17, 26, 41
active layer system for OPV, 72 Conjugative polymers, 63
hybrid photovoltaic cells, 72–73 Copolymers, 73, 74
morphology of, 72, 75 Copper indium diselenide (CuInSe2), 86, 229–231
nonfullerene-based, 73 Copper indium disulfide (CuInS2), 227–229
p-type perylene diimide in, 73 Copper indium gallium diselenide (CIGSe), 232–235

265
266 Index

Copper indium gallium disulfide (CIGS), 231–232 types of, 101


Copper iron tin sulfide and related materials, 242–245 quantum dot DSSC, 104–105
Copper sulfide, 220–222 quasi-solid state DSSC, 103–104
Copper-zinc-tin chalcogenides (CZTSSe), 235–242 solid-state DSSC, 101–103
nanocrystals–based approach, 235–236
nanoparticles, 240–241
power conversion efficiency of, 240 E
processing of, 235
Electrochemical aspect, of hydrogen, 164–167
synthesis of, 237–239
Electrochemical cell, 56, 132; see also
Counterelectrodes, 12, 100–101, 243
Photoelectrochemical (PEC) cells
CSP, see Concentrated solar power (CSP)
Electrochemical photovoltaic cells, 31–32
Cu2CoSnS4, 243, 244
Electrodes
Cu2FeSnS4, 243
doped semiconductors, 98–99
Cu2(FexZn1–x)SnS4 nanocrystals, 243–244
modified semiconductors, 97
Cu2MnSnS4, synthesis of, 244
naïve semiconductors, 95–97
Cyanine dyes, 62
other nanomaterials, 99–100
Electrolysis, of water, 143, 148–149
D solarothermal, 150
thermal, 150–151
Dendrimers, 63, 65
Electrolytes, component in DSSCs, 11, 93
Depletion layer, 21, 25
liquid, 11, 93–94
Differential cell, 117, 118
other, 94–95
Donor-bridge-acceptor (D-B-A) system, 193, 194
polymer gel, 94
Doped semiconductors, 17, 98–99; see also
quasi-solid-state, 11
Semiconductors
solid, 11
DSPC, see Dye-sensitized photoelectrochemical cell
Energy, 1; see also Solar energy
(DSPC)
crisis, 2
DSPECs, see Dye-sensitized photoelectrosynthesis cells
global, use and production, 187–188
(DSPECs)
renewable, 2–3, 7, 86, 140–143, 187, 208
DSSC, see Dye-sensitized solar cell (DSSC)
Energy Conversion Devices, 166
Dyes, 62–63, 91–92
Engineered solid solutions, 159
photogalvanic cell and, 119–124
Excitons, 13, 14
quantum dots (QDs) and, 157–159
diffusion, 58–59
Dye-sensitized photoelectrochemical cell (DSPC), 255
exciton diffusion length, 255
Dye-sensitized photoelectrosynthesis cells (DSPECs),
39, 41
Dye-sensitized solar cells (DSSCs), 10, 85, 195 F
applications of, 105
basic parts of, 10–11 Fe(II)-β-diketonate/thionine, 123
characterization, 88 Fermi level (Ef ), 18, 19, 25, 164
components of, 89 definition, 17
counterelectrode, 100–101 in intrinsic semiconductor, 17, 22
electrodes Fe-thionine photogalvanic cell, 122
doped semiconductors, 98–99 Fill factor (FF), 31, 61, 66, 75, 88
modified semiconductors, 97 First-generation solar cells, 86, 253, 254
naïve semiconductors, 95–97 Formaldehyde, 13
other nanomaterials, 99–100 Formamidinium lead iodide (FAPbI3), 260
electrolytes, 11, 93–95 Formic acid, 208
liquid, 93–94 Fossil fuels, 2, 4, 5, 143, 147–148, 155, 187, 219
other, 94–95 Fuel cells, 14, 168, 208
polymer gel, 94 hydrocarbon waste of high-temperature, 153–154
fabrication of, 101 hydrogen and, 141, 142, 143
four energy levels, 11 Fuel production, catalysis for CO2 reduction and,
heterojunction three-dimensional fabrication of, 86 208–213
incident photon-to-current conversion efficiency, 89 Fullerene, 15, 35, 63, 64, 70, 73, 74, 259
nanocrystalline semiconductor working electrode, β-Functionalized porphyrin sensitizers, 32
95–100
photo-to-electric conversion efficiency of, 99–100 G
potential of, 100
principle, 87 Graphene, 67–68, 166–167
sensitizer, 89 Graphene oxide (GO), 44, 47, 76, 212
dyes, 91–92 Graphene quantum dots (GQDs), 104–105
metal complexes, 89–91 Gratzel, Michal, 10
natural pigments, 93 Gratzel cell, see Dye-sensitized solar cell (DSSC)
Index 267

H Hydrogenases
catalytic action of, 206
Heterocyclic dyes, 124 definition, 204
Heterogeneous artificial photosynthesis systems, 195–196 proton reduction catalyst by, 204–206
Heterogeneous photocatalysis, 26 Hydrogen generation, 140, 142–144
High-resolution transmission electron microscopy light-driven catalysis for, 206–207
(HRTEM), 66 photochemical, 144–147
High-temperature fuel cells, hydrocarbon waste of, resources for, 142–143
153–154 technology for, 147–152
Homogeneous artificial photosynthesis systems, 191–195 biological production, 148
Homogeneous photocatalysis, 26 concentrating solar thermal power, 149–150
Homogenous water oxidation, 201 electrolysis of water, 148–149
Honda FCX concept car, 140–141 Kvaerner carbon black and hydrogen process, 148
HRTEM, see High-resolution transmission electron photocatalytic and photoelectrocatalytic, 151–152
microscopy (HRTEM) thermal electrolysis of water, 150–151
Hybrid solar cells, 14–15, 77, 255–258
dye-sensitized solar cells, 257
nanoscale structures, 255–256 I
organic material, 255 IEA, see International Energy Agency (IEA)
photoactive layer, 256 Incident photon-to-current conversion efficiency (IPCE),
Hybrid tandem photovoltaic cell, 76–77 61, 89, 145–146, 257
Hydrocarbon fuels, 43–44, 147, 211 Infrared (IR) radiation, 7, 59
Hydrogen, 12, 13, 139–141 Inorganic semiconductors, 14, 56, 115, 255, 256
biological production, 148 Insulator, electrical conductivity of, 17
as by-product of other chemical processes, 152–154 Internal combustion engines (ICEs), 168
chlor-alkali plants, 153 Internal reforming, 153–154
high-temperature fuel cells, hydrocarbon waste of, International Energy Agency (IEA), 2, 143
153–154 Intrinsic semiconductor, 17–18
oil refinery, catalytic reforming of, 153 Inverted organic photovoltaic cells, 77–78
waste biomass of wine industries, 154 Inverted structure, 77
chemical fuel, 141 Inverted tandem solar cell, 77
concentrating solar thermal power, 149–150 IPCE, see Incident photon-to-current efficiency (IPCE)
eco-friendly production of, 140 Iridium-based catalyst, 202–203
economy, 168 Iron-iron hydrogenases, 204
electrochemical aspect, 164–167 Iron sulfide, 222–223
electrolysis of water, 148–149 Iron–thionine photogalvanic cell, 115, 120–121
as energy carrier, 144
evolution and catalytic proton reduction and, 204–207
fuel cells, 140, 141, 143 K
generation, technology for, 147–154 Kvaerner carbon black and hydrogen process, 148
as key solution, 167–168
Kvaerner carbon black and hydrogen process, 148
photocatalysts, modification of, 157 L
composite photocatalyst, 163 Langley (L), 8
engineered solid solutions, 159 Lanthanides, 66
photocatalyst, nanostructuring of, 162–163 Lead sulfide, 223–225
photoelectrochemical water splitting, 163–164 LHE, see Light-harvesting efficiency (LHE)
photosensitization, 157–159 Light absorption (LA), 46, 160, 161, 188, 192, 196–197,
reducing bandgaps through doping, 159 209
surface plasmon resonance, plasmonic Light-driven catalytic water splitting, 199
nanostructures with, 161–162 Light-harvesting efficiency (LHE), 146
water splitting, macromolecular systems for, Liquid crystals, 65–66
159–161 Liquid electrolytes, 11, 93–94
photocatalytic and photoelectrocatalytic hydrogen Liquid-junction PEC cell, 31
production, 151–152 Liquid-junction solar cells, see Photogalvanic cells
photocatalytic hydrogen generation, 155–156
photocatalytic water splitting, mechanisms of, 156–157
photochemical hydrogen generation, 144–147 M
photogeneration of, 13 Macromolecular systems for, water splitting, 159–161
as renewable energy, 140 Manganese-molybdenum-diethyldithiocarbamate
as safe energy source, 142 complex, 128
as sustainable and clean energy source, 142 MEG, see Multiple exciton generation (MEG)
as synthetic fuel, 13 Metal-based electrolytes, 95
thermal electrolysis of water, 150–151 Metal complexes, 89–91
268 Index

free dyes, 91–92 Oil shocks, 2


other metal complexes, 90–91 Oligomers, 63
Ru metal complex dyes, 90 Open-circuit voltage (Voc), 38, 60, 64–65, 68, 73, 88, 122
Metal-free organic dyes, 92 Operation conditions, for reduction of CO2, 181–182
Metal oxide photocatalysts, 179 pressure effect, 181
Metal-to-ligand charge-transfer (MLCT) transitions, 35, reactor-type effect, 182
196, 208 reductant effect, 181
Methylammonium lead bromide (MAPbBr3), 260 temperature effect, 181
Microbial photoelectrochemical cell (MPC), 258 Organic cell, electrochemical cell vs., 56
Mn-, Co-, and Fe-based catalysts, 203–204 Organic chromophore dyes, 91–92
Modified semiconductors, 97 Organic dyes, 62–63
Molecular donor–acceptor systems, charge separation by, Organic light-emitting devices (OLEDs), 58
197–198 Organic materials, for photovoltaic applications, 61–62
Molecular light-harvesting systems, 193 Organic photovoltaic cells, 13–14, 55–78
Mott–Schottky equation, 22 advantages of, 56
MPC, see Microbial photoelectrochemical cell (MPC) basic processes, 58–60
Multiple exciton generation (MEG), 224 charge collection, 60
charge separation, 59
N charge transfer, 59–60
exciton diffusion, 58–59
Naïve semiconductors, 95–97 photon absorption, 58
Nanocrystalline semiconductor working electrode, 95–100 characteristics, 60–61
doped semiconductors, 98–99 history, 57–58
modified semiconductors, 97 hybrid tandem photovoltaic cell, 76–77
naïve semiconductors, 95–97 introduction, 55–57
other nanomaterials, 99–100 inverted, 77–78
Nanomaterials for solar energy, 219–220 inverted tandem solar cell, 77
binary materials, 220 materials, 61–65
copper sulfide, 220–222 dyes, 62–63
iron sulfide, 222–223 liquid crystals, 65–66
lead sulfide, 223–225 other materials, 66–68
tin sulfide, 225–226 pigments, 62
copper-zinc-tin chalcogenides, 235–242 small molecules, oligomers, polymers, and
copper iron tin sulfide and related materials, 242–245 dendrimers, 63–65
I–III–VI materials, 226–227 MoS2 with, 67
copper indium diselenide, 229–231 tandem solar cells, 75–76
copper indium disulfide, 227–229 types of, 68–75
copper indium gallium diselenide, 232–235 bilayer, 69–71
copper indium gallium disulfide, 231–232 heterojunctions, 71–75
Natural dye pigments, 92 single layer, 68–69
Natural photosynthesis, 9 Organic semiconductors, 56, 57, 70
to artificial photosynthesis, 188–190 Organisation for Economic Co-operation and
at molecular level, 193 Development (OECD), 142
n-Doped semiconductor, 17, 19, 165; see also p-doped Oxide photocatalysts, 163
semiconductor
effect of pH on potential drop in, 24 P
effect on band bending and Fermi level on, 25
photoexcitation of, 19 PbS quantum dots, 224–225
redox couple in vacuum, 20–21 PCSNPs, see Plasmonic core–shell nanoparticles
reverse band-bending, 19 (PCSNPs)
space charge layer in, 21 p-Doped semiconductor, 18, 19; see also n-doped
in vaccum, 23 semiconductor
n-DSSCs, 10 band bending in, 19, 20
Nevada Policy Research Institute (NPRI), 3 electrolyte with a redox couple, 23
n-GaAs electrodes, 32, 34 redox couple in vacuum, 22
NiO photocathode–based dye-sensitized solar cell, 99 reverse band-bending, 19, 20
Nonoxide semiconductor photocatalysts, 179–181 space charge layer in, 22
n-p-Type cells, 20 p-DSSC, 10
PEBFC, see Photoelectrochemical biofuel cell (PEBFC)
O Perovskite solar cells, 14, 15, 259–260
Phenozine dyes, 123
OECD, see Organisation for Economic Co-operation and Photocatalysis, 26–28, 175
Development (OECD) definition, 26
Oil refinery, catalytic reforming of, 153 example, 145
Index 269

heterogeneous, 26 photochemical determinants of, 124–125


homogeneous, 26 of rubidium anion, 126–127
various probabilities of reactions, 26–27 requirements for efficient, 117–119
Photocatalysts, 145 theory of the operation, 120
composite, 163 Photogalvanic device functions, 12, 115, 132
modification of, 157 Photogalvanovoltaic (PGV) cell, 130
composite photocatalyst, 163 Photoinitiated electron collection (PEC) process, 160
engineered solid solutions, 159 Photon absorption, 58
photocatalyst, nanostructuring of, 162–163 Photonic flux density, 9
photoelectrochemical water splitting, 163–164 Photoredox process, 116
photosensitization, 157–159 Photosensitization, 34, 124, 157–159
reducing bandgaps through doping, 159 Photosynthesis, 9–10, 119, 145,155, 163, 258
surface plasmon resonance, plasmonic artificial, see Artificial photosynthesis
nanostructures with, 161–162 natural, see Natural photosynthesis
water splitting, macromolecular systems for, 159–161 reactions, 188–189
nanostructuring of, 162–163 Photosystem I (PS I), 14, 188–189, 190, 258
for reduction of CO2, 177–181 Photosystem II (PS II), 14, 188–189, 191, 198
mechanism, 174–176 Photovoltaic cells, 3, 13–14, 55–56, 57, 130–131
metal oxide photocatalysts, 179 bilayer organic, 69–71, 72
nonoxide semiconductor photocatalysts, 179–181 electrochemical, 31–32
TiO2 and related titanium-containing solids, 177–179 hybrid tandem, 76–77
Photocatalytic cells, 30, 39, 42–47 inverted organic, 77–78
Photocatalytic hydrogen generation, 47, 155–156 organic, see Organic photovoltaic cells
Photocatalytic water splitting silicon-based, 56
analogy of photosynthesis with, 145 single-layer organic, 68–69
for hydrogen production, 42, 47, 156–157 Photovoltaic (PV) effect, 220
mechanisms of, 156–157 Photovoltaic/photoelectrochemical (PV/PEC), 258
Photochemical conversion modes, 9–15 Pigments, 62
Photochemical hydrogen generation, 144–147 natural, 92
Photochemical reaction, 117–119 Planar donor–acceptor heterojunction, 69
Photoconductivity, 18, 57, 62, 63, 257 Planar donor–acceptor hydrogen, 14
Photoconversion efficiency (η), 42, 47, 146–147 Plasmonic core–shell nanoparticles (PCSNPs), 254–255
Photocurrent enhancement, 75, 253–254 Plasmonic nanostructures, 161–162
Photoelectrochemical biofuel cell (PEBFC), 257 Plasmonic solar cells, 15, 253–255
Photoelectrochemical (PEC) cells, 12, 196 Plasmon resonance energy transfer (PRET), 161
applications of, 12 Plastic solar cell (PSC), 15, 63, 260
classification, 30–31 Poly(adipic acid pentaerythritol ester) (PAAPE), 104
photocatalytic cells, 42–47 Poly(3-hexylthiophene) (P3HT), 71, 78
photoelectrosynthetic cells, 30–31, 39–42 Poly(vinyl carbazole) (PVK), photoconductivity of, 63
regenerative PEC solar cells, 30, 31–39 Polymer gel electrolytes, 94, 103
commercial use of, 29 Polymers, 14, 15, 35, 57, 62, 63–65, 71–72, 73
definition, 29 Polymer solar cell, 260
energy production by, 29 Polythiophene, 63, 68
semiconductor–liquid-junction-based, 30 Porphyrin quinine (P−Q) system, 197–198
as storage cells, 47–48 Power conversion efficiency (PCE, η), 14, 32, 38, 57, 61,
titanium dioxide in, 12 70, 73, 105, 260
Photoelectrochemical (PEC) reactions, 163 Power point, 88
Photoelectrochemical water splitting, 163–164 PRINT (pattern replication in nonwetting templates), 58
Photoelectrochemistry, 12, 24–26, 122, 131 Proton reduction catalyst, by hydrogenases, 204–206
Photoelectrolysis, of water, 129 PSt-bpy-Ru complex, 32–33
Photoelectrolytic solar cells, 39–42 Pt electrode, 100, 101
Photoelectrosynthetic cells, 30–31, 39–42 Push–pull conjugated copolymers, synthesis of, 73
Photogalvanic cells, 12, 115, 116, 124, 128
cell efficiency, 117 Q
complexes, 124–129
definition, 115 Quantum dot dye-sensitized solar cells (QD-DSSC),
dyes, 119–124 104–105
at electrode, 11 Quantum dots, 3, 11, 261
examples, 125–126 CdS, 104
functioning of, 120 colloidal, 76, 77
introduction, 115–117 dyes and, 157–159
miscellaneous, 129–132 graphene, 104–105
operation PbS, 224–225
construction, 128–129 solar cell, 261
270 Index

Quantum efficiency (QE), 61, 145, 147 advantages, 3–4


Quasi-solid-state DSSC (QSS-DSSC), 11, 103–104 conversion efficiency, 39–40
Quasi-solid-state electrolytes, 11 disadvantages, 4
future, 4–5
R in generating electricity, 2–3, 4
nanomaterials for, 219
Regenerative PEC solar cells, 30, 31–39 copper indium diselenide, 229–231
Renewable energy resources, 2–3, 7, 86, 140–143, 187, 208 copper indium disulfide, 227–229
Reverse water gas shift (RWGS) reaction, 208 copper indium gallium diselenide, 232–235
Ru(II) complex dyes, 90, 91, 92, 102, 202 copper indium gallium disulfide, 231–232
Ruthenium-based catalysts, 200–202 copper iron tin sulfide and related materials,
242–245
S copper sulfide, 220–222
copper-zinc-tin chalcogenides, 235–242
Schottky-type cells, 20 iron sulfide, 222–223
Schottky-type junction, 21 lead sulfide, 223–225
Second-generation solar cells, 86, 253, 254 tin sulfide, 225–226
Selenium, photoconductive effect in, 57 scenario, 7–9
Semiconductors for water breaking, 145
electrochemistry of, 17–24 Solar fuel, 9, 10, 195
electron distribution, 17, 18 Solar influx, 8
Fermi level, 17, 18 Solar radiation, 1, 55, 85
material classification, 17, 18 annual mean of, 8
n-doped semiconductors, see n-doped semiconductors conversion into electrical energy, 58
p-doped semiconductor, see p-doped semiconductors charge collection, 60
Sensitization, 119, 157 charge separation, 59
Sensitizer, 89 charge transfer, 59–60
dyes, 91–92 exciton diffusion, 58–59
metal complexes, 89–91 photon absorption, 58
other metal complexes, 90–91 unit of, 8
Ru metal complex dyes, 90 Solar spectrum, 7–8, 194, 196
metal complex free dyes, 91–92 Solar-to-electricity conversion efficiency, 33, 92
natural pigments, 93 Solar-to-hydrogen (STH) conversion efficiency, 145, 146
Short-circuit current (isc), 60–61 Solid-state dye-sensitized solar cell, 101–103, 257
Silicon-based photovoltaic cells, 56 Solid-state electrolytes, 11, 91, 93, 103
Silicon solar cells, 3, 226 Solid-state photogalvanic dye-sensitized solar cells, 124
Silver nanoprisms (AgNPs), 255 Space charge layer, see Depletion layer
Single-layer organic photovoltaic cells, 68–69 Spectral distribution, of sunlight, 8–9
Single-site photocatalysts, 177–178 Spray pyrolysis technique, 228
Single-walled carbon nanotube (SWNT), in bulk Steam reforming, 208
heterojunction, 75 Stille polymerization reaction, 72
Solar cells, 1, 3, 86 Successive ionic layer adsorption and a reaction (SILAR)
biohybrid, 14, 258–259 process, 37
cascade heterojunction organic, 75 Sunlight, 3, 7
characteristics, 60–61 diffusion of, 8
dye-sensitized, see Dye-sensitized solar cells (DSSCs) spectral distribution of, 8–9
first-generation, 86, 253, 254 Surface plasmon resonance (SPR), 161–162
hybrid, see Hybrid solar cells
inverted tandem, 77 T
perovskite, 14, 15, 259–260
photoelectrolytic, 39–42 Tandem solar cells, 10, 75–76
plasmonic, 15, 253–255 colloidal quantum dots and, 76
plastic, 15, 63, 260 interlayer, 76
polymer, 260 multilayer, 76
quantum dot dye-sensitized solar cells, 104–105 single-layer, 75–76
regenerative PEC, 30, 31–39 Thermal electrolysis, of water, 150–151
second-generation, 86, 253, 254 three-step cycle, 151
silicon, 3, 226 two-step cycles, 150–151
solid-state dye-sensitized, 257 Thiazine dyes, 120
tandem, see Tandem solar cells Thin-film solar cells, 253
third-generation, 86, 253, 254 Thionine-EDTA system, for solar energy conversion, 122
Solar constant, 7 III–V materials, 3
Solar electricity, 55–56, 220 Third-generation solar cells, 86, 253, 254
Solar energy, 1, 2, 85–86, 219–220 Tin sulfide, 225–226
Index 271

Titanium dioxide (TiO2), 10, 12, 96, 156 Water


coated SnO2 hybrid nanorod, 99 electrolysis of, 148–149
Fe-doped titania, 98 splitting
MgO-coated, 97 by Ir complex, 202
nanofibers, 100 light-driven catalytic, 199
nitrogen-doped, 98 macromolecular systems for, 159–161
photocatalytic activity of, 177 phenomena, 156–157
photoreduction of CO2 by using, 177 photoelectrochemical, 163–164
titanium-containing solids and, 177–179 thermal electrolysis of, 150–151
for water splitting, 156 Water electrolysis cell, 145, 149
Trans-genes, 205 Water oxidation, 44, 189, 199
Transition-metal catalysts, 194, 210 catalytic, 40, 41, 42, 199–204
Ruthenium-based catalysts, 200–202
U Iridium-based catalyst, 202–203
Mn-, Co-, and Fe-based catalysts, 203–204
Ultraviolet (UV) radiation, 7, 8, 67, 178 homogenous, 201
Wine industries, waste biomass of, 154
V
Valence band (VB), 11, 17, 22, 26, 157 X
Vanadium oxides (VOx), 77
Visible (VIS) radiation, 7 Xanthene dyes, 123

W Z
Waste biomass, of wine industries, 154 ZnO nanorod, 99, 104

You might also like