(Series On Photoconversion of Solar Energy) Mary D Archer, Martin A Green, Mary D Archer, Martin Green - Clean Electricity From Photovoltaics (2nd Edition) - Imperial College Press (2014) PDF
(Series On Photoconversion of Solar Energy) Mary D Archer, Martin A Green, Mary D Archer, Martin Green - Clean Electricity From Photovoltaics (2nd Edition) - Imperial College Press (2014) PDF
FROM PHOTOVOLTAICS
2nd Edition
CLEAN ELECTRICITY
FROM PHOTOVOLTAICS
2nd Edition
Editors
Mary D Archer
Imperial College, UK
Martin A Green
University of New South Wales, Australia
Distributed by
World Scientific Publishing Co. Pte. Ltd.
5 Toh Tuck Link, Singapore 596224
USA office: 27 Warren Street, Suite 401-402, Hackensack, NJ 07601
UK office: 57 Shelton Street, Covent Garden, London WC2H 9HE
For photocopying of material in this volume, please pay a copying fee through the Copyright Clearance
Center, Inc., 222 Rosewood Drive, Danvers, MA 01923, USA. In this case permission to photocopy
is not required from the publisher.
ISBN 978-1-84816-767-4
Printed in Singapore
to
v
May 2, 2013 14:6 BC: 8831 - Probability and Statistical Theory PST˙ws
CONTENTS
vii
September 17, 2014 11:29 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-fm page viii
viii Contents
Contents ix
x Contents
Contents xi
Appendices 663
Index 671
May 2, 2013 14:6 BC: 8831 - Probability and Statistical Theory PST˙ws
Mary Archer read chemistry at Oxford University and took her PhD from Imperial
College London, in 1968. From 1968 to 1972, she did post-doctoral work in elec-
trochemistry with Dr John Albery at Oxford, and she then spent four years at The
Royal Institution in London, working with Lord Porter (then Sir George Porter)
on photoelectrochemical methods of solar energy conversion. She taught chem-
istry at Cambridge University from 1976 to 1986. From 1991 to 1999, she was a
Visiting Professor in the Department of Biochemistry at Imperial College London,
and from 1999 to 2002, she held a Visiting Professorship at ICCEPT (Imperial
College Centre for Energy Policy and Technology). She is President of the UK
Solar Energy Society and the National Energy Foundation and a Companion of the
Energy Institute. She was awarded the Melchett Medal of the Energy Institute in
2002 and the Eva Philbin Award of the Institute of Chemistry of Ireland in 2007.
In 2012, she was appointed a Dame Commander of the British Empire for services
to the UK National Health Service.
Christophe Ballif received his MSc and PhD degrees in physics from the Fed-
eral Polytechnic School of Lusanne, Switzerland, in 1994 and 1998, respectively,
focusing on novel photovoltaic materials. Following post-doctoral research at the
National Renewable Energy Laboratory, Golden, USA, on CIGS and CdTe solar
cells, he moved to the Fraunhofer Institute for Solar Energy Systems, Freiburg,
Germany, where he focused on crystalline silicon photovoltaics until 2003. He
then joined the Swiss Federal Laboratories for Materials Testing and Research,
Thun, Switzerland, before becoming a full professor at the Institute of Micro-
engineering, University of Neuchâtel, Switzerland, in 2004. In 2009, the Institute
of Microengineering was transferred to EPFL. He is the Director of the Photo-
voltaics and Thin-Film Electronics Laboratory in the Institute, and since 2013 he
has also been the Director of the PV-Centre of the Swiss Centre for Electronics
and Microtechnology (CSEM), Neuchâtel. His research interests include thin-film
silicon, high-efficiency heterojunction crystalline cells, module technology, con-
tributing to technology transfer, and industrialisation of novel devices.
Dieter Bonnet was born in Stuttgart, Germany, in 1937 and obtained his PhD on
the photoelectric properties of organic materials at Frankfurt University in 1963.
In 1965, he joined Battelle Institute in Frankfurt, where in 1968 he started work on
thin-film solar cells based on II–VI compounds, including CdTe. In 1970, he made
xiii
September 17, 2014 11:29 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-fm page xiv
the world’s first CdTe/CdS thin-film solar cell in the presently known configuration.
After a period working on other solar cell materials, he resumed work on CdTe
technology in 1990, and initiated the successful EU (European Union) projects
EUROCAD in 1992 and the CdTe interest group SOLARPACT in 2005. In 1993, he
co-founded the pioneer company ANTEC GmbH to manufacture CdTe modules.
In May 2001, the road in which ANTEC is located in Arnstadt, Germany, was
named Dr. Bonnet Weg in honour of his accomplishments, and he was awarded
the Becquerel Prize of the European Commission for outstanding achievements in
photovoltaics in 2006. He retired in 2001 and is now an independent consultant.
Dan Credgington received his MSci in Natural Sciences from the University of
Cambridge in 2004, and was awarded the Herchel Smith scholarship to study at Har-
vard University. In 2010, he obtained his PhD for work on the nanoscale microscopy
and lithography of conjugated molecules from University College, London, under
the supervision of Professor Franco Cacialli, and went on to conduct post-doctoral
research on organic solar cells with Professor James Durrant at Imperial College
London. He is currently a post-doctoral researcher in the group of Professor Sir
Richard Friend at the Cavendish Laboratory of the University of Cambridge, where
his interests lie in the study of recombination processes in organic light emitting
diodes and solar cells, and hybrid organic/inorganic technologies.
Arnulf Jäger-Waldau received his Dr. rer. nat. from the Physics Department of
the University of Konstanz, Germany, in 1993. He has worked in the field of
material research for solar cells since 1987 and holds patents on semiconductor
material deposition for thin-film solar cells and solar module design. In 1994 and
1995 he worked as a post-doctoral JSPS fellow at Shinshu University, Nagano,
Japan, before joining the Hahn–Meitner Institute Berlin in 1996. Since 2001 he
has been a Scientific Officer and Senior Scientist at the Renewable Energy Unit,
Institute for Energy and Transport of the European Commission’s Joint Research
Centre, where he works on the assessment of renewable energy technologies, the
effectiveness of their implementation, their integration into energy infrastructures
and the role of renewable energy for climate change mitigation. Among other
roles, he has been the Technical Chairman of the European Photovoltaic Solar
Energy Conference (EUPVSEC) since 2011, and he was a Lead Author for Solar
Energy of the Special Report of the IPCC on Renewable Energy and Climate
Change Mitigation. He also serves as a member of the Executive Committee of
the European Materials Research Society, the Academic Advisory Board of the
Chinese Trina State Key Laboratory for Photovoltaics, the International Advisory
Board of the Warsaw University Photovoltaic Centre and the Scientific Advisory
Board of the Solar Research Centre of the Bulgarian Academy of Science, and he
is Vice-Chairman of the Academic Committee of the Asian Photovoltaic Industry
Association (APVIA).
Antonio Luque obtained his Doctor of Engineering degree from the Polytechnic
University of Madrid in 1967. Today he is also Doctor Honoris Causa of three other
Spanish universities. In 1969, he joined the university staff and founded its Semi-
conductor Laboratory. In 1979, this centre became the Institute of Solar Energy that
he leads at present. In 1981, he founded the company Isofotón to manufacture the
bifacial cells he had invented, and he chaired its board until 1990. He has written
September 17, 2014 11:29 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-fm page xviii
some 300 papers and registered around 20 patents, of which some are in exploita-
tion. He has won several scientific awards, among which are the Spanish National
Prize for Technology in 1987 and 2003, the King Jaime I Prize for environmental
protection, the Becquerel Prize awarded by the European Commission for PV in
1992, the IEEE William Cherry Award in 1996 and the SolarWorld Einstein Award
in 2008. Among other distinctions, he is a member of the Royal Academy of Engi-
neering, and of the Russian Academy of Sciences. He has also been a member
of the Advisory Council for Science and Technology, which advises the Spanish
Prime Minister.
Ignacio Luque-Heredia is CEO of BSQ Solar, a high concentration photovoltaics
(HCPV) manufacturing company that he co-founded in 2009. He received his MSc
and PhD degrees in electrical engineering from Polytechnic University of Madrid
in 1995 and 2010 respectively. He was co-founder in 1995 of the company Inspira,
also operating in the CPV field, which was acquired by the Silicon Valley leading
CPV manufacturer Solfocus in 2007. From 2007 to 2009 he was CTO for Solfocus
Europe. Leading Inspira’s and BSQ Solar’s engineering, he has participated in 23
collaborative projects in the field of CPV, 11 of them funded by the European
Commission. These range from the EUCLIDES project in 1995, which resulted
in the biggest CPV plant of its time, to the most recent NGCPV project, on 3rd
generation photovoltaic devices and their integration in CPV technologies. He
holds four patents and has led the deployment of CPV pilot systems and large-
scale plants, as well as several technology transfer programs in the USA, China,
India, Japan, Australia, MENA, Brazil, Mexico and Europe. He is a member of
the Scientific Committee of the International Conference on CPV Systems, and
of Work Group 7 for the development of CPV standards in Technical Committee
82 for Solar Photovoltaic Energy Systems of the International Electrotechnical
Commission. He is also a Senior Member of the IEEE.
Jenny Nelson is a Professor of Physics at Imperial College London, where she
has researched novel types of solar cell since 1989. Her current research focuses
on photovoltaic energy conversion using molecular materials, characterisation of
the charge transport, charge separation and morphological properties of molecu-
lar semiconductors, the theory of charge transport in organic semiconductors and
modelling of photovoltaic device behaviour. She has published over 200 papers on
photovoltaic materials and devices, and a book on the physics of solar cells.
Nicola Pearsall is a Professor at Northumbria University, where she leads their
photovoltaic research activities. She holds a degree in physics from the University
of Manchester Institute of Science and Technology and obtained her PhD from
Cranfield Institute of Technology for her research on indium phosphide solar cells
September 17, 2014 11:29 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-fm page xix
for satellite applications. She has been involved in research in photovoltaics for 30
years, and her current interests relate to the performance and implementation of
photovoltaic systems. She is a member of the Steering Committee of the European
Photovoltaic Technology Platform and has contributed to the development of their
Strategic Research Agenda, as well as working with the Solar Europe Industry
Initiative.
Uwe Rau received his PhD in physics in 1991 from the University of Tübingen,
Germany, for his work on temporal and spatial structure formation in the low-
temperature electronic transport of bulk semiconductors. From 1991 to 1994, he
worked at the Max Planck Institute for Solid State Research, Stuttgart, on Schottky
contacts, semiconductor heterojunctions and silicon solar cells. From 1994 to 1997,
he worked at the University of Bayreuth, Germany, on electrical characterisation
and simulation of Si and CuInSe2 solar cells. In 1997, he joined the Institute for
Physical Electronics at the University of Stuttgart, where he became leader of
the Device Analysis Group. His research interests centre on transport phenomena,
especially electrical transport in solar cell heterojunction devices and interface and
bulk defects in semiconductors. He has authored or co-authored more than 100
scientific publications.
Hans-Werner Schock received his diploma in electrical engineering in 1974, and
doctoral degree in electrical engineering in 1986, from the University of Stuttgart’s
Faculty of Electrical Engineering. Since the early 1970s, he has worked on the
development of polycrystalline II–VI and I–III–VI2 compound semiconductor thin-
film solar cells, taking the development of chalcogenide solar cells from research
to pilot fabrication. A series of successful research projects on thin-film solar cells
under his guidance resulted in several production lines for thin film solar cells
in Europe. From 2004 to 2012, he was director of the Institute of Technology at
the Helmholtz Zentrum Berlin for Materials and Energy and he is an Honorary
Professor at the Technical University Berlin. He received the Becquerel Prize of
the European Commission in 2010 for his achievements in the development of
thin-film solar cells. At present he is a consultant in the field of photovoltaics.
He is the author or co-author of more than 300 contributions in books, scientific
journals and conference proceedings.
Masafumi Yamaguchi is a Professor of the Toyota Technological Institute. He
received his BS and PhD degrees from Hokkaido University, Japan, in 1968 and
1978, respectively. In 1968, he joined the NTT Electrical Communications Labo-
ratories, working on radiation damage to Si and III–V compounds, ZnSe blue-
light-emitting diodes and III–V compound solar cells. In 1983, he discovered
the superior radiation-resistance of InP materials and solar cells, thereby showing
September 17, 2014 11:29 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-fm page xx
the great potential of InP cells for space applications. His group also developed
high-efficiency InP, GaAs-on-Si, and AlGaAs/GaAs tandem cells by proposing
a double-hetero structure tunnel junction for realising a high performance and
stable multijunction cell interconnection in 1987. As Japanese team leader of the
EU–Japan Collaborative Research on Concentrator Photovoltaics, he contributed to
the attainment of InGaP/GaAs/InGaAs 3-junction cells with efficiencies of 44.4%
at 302 Suns of AM1.5D and 37.9% at AM1.5G. He is also the project leader
of the Next Generation High Performance Photovoltaics Research and Develop-
ment Project of the Japanese New Energy Development and Industrial Technology
Development Organisation, and the Research Supervisor in the Research Area
of the Creative Clean Energy Generation using Solar Energy programme of the
Japan Science and Technology Agency. He has published more than 300 origi-
nal papers and received numerous awards, such as the Becquerel Prize from the
European Commission in 2004 and the William Cherry Award from the IEEE
in 2008 for outstanding contributions to science and technology development of
high-efficiency photovoltaics.
September 17, 2014 11:29 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-fm page xxi
Since the dawn of history, man has been fascinated by the Sun, the provider of
the light and warmth that sustains life on Earth. In pre-industrial times, our major
sources of energy — wood, wind and water power — derived from solar energy.
The subsequent discovery and massive exploitation of fossil fuels laid down in
the Earth’s crust by early aeons of photosynthetic activity have conditioned the
developed world to be dependent on convenient, readily available energy. But we
are living on our energy capital. The Earth’s reserves of coal, oil and gas are
finite and likely to become resource-depleted in the course of this century. A sense
of living on borrowed time was therefore appropriate even before concerns about
global climate change, sustainability and energy security combined to raise interest
in renewable energy to its current encouraging level.
This book is the first in a series of four multi-authorial works on the pho-
toconversion of solar energy. It was created from my long-held conviction that,
despite slow starts and setbacks, solar energy — broadly defined to encompass
other renewable energy forms that derive from solar — will become the Earth’s
major energy source within this century. The Sun is a source of both radiant heat
and light, and techniques for using solar energy correspondingly divide into ther-
mal methods (solar power towers, water heaters and so on) and photoconversion
(sometimes called direct) methods. Photoconversion is the subject of this book
series. A photoconverter is a device that converts sunlight (or any other source of
light) into a useful form of energy, usually electrical power or a chemical fuel, in
a process that relies, not on a raised temperature, but on the selective excitation
of molecules or electrons in a light-absorbing material and their subsequent de-
excitation in a way that produces energy in a useful form. Volume I covers the most
developed of the man photoconversion devices, photovoltaic (PV) cells, which are
solid-state semiconductor devices that produce electrical power on illumination.
Volume II will cover the natural photoconversion system of photosynthesis, the
potential of biomass as an energy source and the global carbon budget. Volume
xxi
September 17, 2014 11:29 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-fm page xxii
III will explore the less developed but exciting possibilities of synthesising arti-
ficial ‘molecule-based’ photoelectrochemical or photochemical photoconverters.
Finally, Volume IV will draw together the common themes of photoconversion and
provide some background material.
The series is intended mainly for senior undergraduates, graduate students and
scientists and technologists working on solar photoconversion. Chapters 1–12 of
this book deal with PV cell design, device physics and the main cell types — crys-
talline and amorphous silicon, cadmium telluride and copper indium diselenide —
as well as more advanced or less developed options such as quantum-well and
thermophotovoltaic cells. These chapters are mainly technical, requiring sound
knowledge of physics, chemistry or materials science for ready understanding.
Chapters 13–18 deal with PV systems, manufacturers, markets and economics and
are accessible without specialist knowledge.
A multi-authorial work owes its very existence to its authors, and my whole-
hearted thanks must go to the twenty-five distinguished individuals, all recognised
authorities in their own fields, who have contributed to this book and patiently
answered my queries during the editing stage. I have also been helped by dis-
cussions about PV with many friends and colleagues, and visits to installations
throughout the world: I have been up Swiss mountains, onto Japanese rooftops and
into the Arizona desert, and thoroughly enjoyed every minute. I am most grateful
to those who have read and commented on various parts of this book or provided
specialist information in advance of publication: Dennis Anderson, Jeffrey and
William Archer, Stephen Feldberg, Martin Green, Eric Lysen, Larry Kazmerski,
Bernard McNelis and Nicola Pearsall. I also warmly thank Alexandra Anghel,
Barrie Clark, Stuart Honan and my PA Jane Williams for editorial assistance, and
Ellen Haigh and John Navas of IC Press and Alan Pui of World Scientific Press for
guiding the book to publication.
For me the sad part of writing this preface is that I must do so in the first
person, for my co-editor Professor Robert Hill died suddenly on 26 November
1999. Bob was the most knowledgeable champion of photovoltaics in the UK, and
his premature death has deprived the British PV community of its cornerstone. He
had drafted his chapter with Nicky Pearsall some months before he died, and the
flow of emails delivering his astute editorial comments on other chapters continued
until the day before his death.
Bob believed unshakeably in the future of PV. Although he knew that system
costs will have to fall by another factor of 2–3 if PV is to become cost-competitive
in major new grid-accessible markets, there are good grounds for believing this is
possible. PV technology is still young, and significant further economies of scale
from larger manufacturing facilities, as well as further advances in the fundamental
September 17, 2014 11:29 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-fm page xxiii
The world of photovoltaics has advanced at a phenomenal pace since the first edition
of this book was published in 2001. Then we were able to report with modest pride
that 200 MWp of PV had been installed worldwide during 1999, taking global
cumulative installed capacity to just over 1 GWp. By the end of 2012, global
installed capacity had edged past the iconic figure of 100 GWp. The final figure for
global installations in 2013 is yet to be determined, but it will add around another
35 GWp, with plausible estimates for cumulative installed capacity by 2020 topping
300 GWp. PV is now, after hydro and wind, the third most important renewable
energy source in terms of installed global capacity.
This prodigious growth has impacted on all sectors of the PV industry. In
1999, much of the small quantity of pure silicon needed by the cell manufacturing
industry came from the moderate resistivity p-type waste material discarded by the
electronics industry. As the PV market grew, a tipping point was reached in 2006
when, for the first time, over half of the world’s supply of polysilicon was used for
PV production. Silicon feedstock for the PV industry is now made by dedicated
plant, and manufacturing capacity has swung sharply away from the USA and
Europe to the lower-cost economies of the Far East.
As the PV market has grown, so the price of cells and modules has dropped.
In their early years, photovoltaic cells were an expensive and exotic novelty for use
in space. In 1955, market leader Hoffman Electronics were offering 2% efficient
silicon cells at a price of $1500/Wp (about 9520/Wp in today’s money). At the
end of 2013, spot prices on the European market were nearly 20,000 times lower
at 0.5–0.7/Wp for silicon modules of efficiencies in the range 15–18%.
xxv
September 17, 2014 11:29 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-fm page xxvi
Between the two editions of this book lie nearly 15 years of scientific discov-
ery, technological advance and market maturation, punctuated by global economic
and financial turmoil, international bickering about the need to mitigate climate
change, national anxieties about energy security, stability and independence, and
political and regulatory instability as governments reconsider their commitment to
renewable energy sources and carbon offset.
The architecture of this edition follows the lines of the first edition, with a
set of chapters about the main PV technologies following an introductory chapter.
Omitted from this edition are thermophotovoltaics, PV for space applications and
electricity storage, while 3-junction III–V concentrator cells of nearly 45% effi-
ciency, CZTS (copper–zinc–tin–sulphide/selenide), dye-sensitised and perovskite
solar cells are notable newcomers to the stage. A further addition is an overview
of the limits to photovoltaic conversion efficiency and how far they can be pushed
beyond single-junction detailed-balance constraints. The book concludes with a
look at PV systems, applications and markets.
We are deeply indebted to all our authors, without whose generosity with their
knowledge and patience with the editorial process this book could not have been
produced. We are particularly grateful to those who have stayed with us from the
first edition and undertaken the painstaking task of radically updating their chapters:
Dieter Bonnet, Antonio Luque, Jenny Nelson (writing about QW solar cells in
the first edition and OPV in the second), Nicola Pearsall, Hans-Werner Schock
and Masafumi Yamaguchi. We warmly welcome newcomer authors Christophe
Ballif, Dan Credgington, Matthieu Despeisse, James Durrant and Michael Grätzel
(who jointly contributed a chapter to Volume III of this series), Ned Ekins-Daukes,
Timothy Gessert, Franz-Josef Haug, Arnulf Jäger–Waldau, Ignacio Luque-Heredia
and Uwe Rau.
The structure of this book series has altered since the preface to the first
edition written: Volume III did not deal with solar fuels as originally planned, but
with photoelectrochemical and nanostructured devices for solar photon conversion;
solar fuels are now planned for the fourth and final volume.
Editing this book series has become a way of life for one of us (MA). She
thanks her family for their tolerance of this enduring preoccupation, her PA Carol
Burling for her skill at the electronic keyboard, James Archer, Richard Friend,
Larry Kazmerski, Bernard McNelis and colleagues at the UK’s annual PVSAT
conferences for much helpful advice and input. MG wishes to thank his partner,
Judy Green, for her ongoing support and Joyce Ho for her assistance on several
fronts during book preparation, as well as the many colleagues and students who
have helped make photovoltaics such a stimulating field of interest.
September 17, 2014 11:29 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-fm page xxvii
CHAPTER 1
1.1 Introduction
Photovoltaic (PV) cells generate electric power when illuminated by sunlight or
artificial light. They are by far the most highly developed of the man-made pho-
toconversion devices. Born of the space age in the 1950s, their earliest terrestrial
applications emerged in the 1970s and they have enjoyed rapid market expansion
since the turn of the millennium. PV technology has a number of advantages over
conventional methods of electricity generation. First and foremost, solar energy
is the world’s major renewable energy resource. PV power can be generated from
the Sun anywhere — in temperate or tropical locations, in urban or rural environ-
ments, in distributed or grid-feeding mode — where there is adequate light. As a
fuel-free distributed resource, PV could in the long run make a major contribution
to national energy security and carbon dioxide abatement. PV is uniquely scalable,
the only energy source that can supply power on a scale of milliwatts to megawatts
from an easily replicated modular technology with excellent economies of scale
in manufacture. A standard 156 × 156 mm crystalline silicon PV cell generates
about 4 peak watts1 (Wp) of DC power and typical PV 60-cell and 72-cell modules
about 240 Wp and 300 Wp respectively. The world’s largest PV generating facility
is currently the Agua Caliente Solar Project in Arizona, which had 251.3 MWp
installed as at November 2013.
1 The power output of a PV cell or module is rated in peak watts (Wp), meaning the power output
at 25◦ C under standard AM1.5 solar radiation of global irradiance 1 kW m−2 . PV ratings in ‘watts’
invariably mean peak watts. To convert from peak watt rating to 24-hour average power output in a
sunny location, divide by a factor of ∼5.
1
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch01 page 2
2 M. D. Archer
PV cells are made of thin semiconductor wafers or films. They contain small
amounts only of (usually non-toxic) materials and, when manufactured in volume,
have modest embedded energy, with energy payback times for silicon PV modules
in Southern Europe of about one year (Fraunhofer, 2012). They possess no moving
parts, generate no emissions, generally require no cooling except when used in
concentrator systems and are silent in operation. PV systems are reliable, easy to
use and long-lived if properly maintained (most commercial modules have lifetime
warranties of 25–30 years, though some balance-of-system components are less
reliable and long-lived than this). Carefully designed, PV arrays are not visually
intrusive, and can indeed add architectural merit to the aesthetic of a built structure.
The PV industry has advanced dramatically in the decade since the first edi-
tion of this book was published. Over the period 2000–2012, solar PV was the
fastest-growing renewable power technology worldwide, and its market growth
has outstripped even the most bullish of forecasts: cumulative global installed PV
capacity in 2000 was only 1.4 GWp, while by the end of 2012 it had just edged past
the milestone of 100 GWp installed (EPIA, 2013). Despite the difficult economic
situation, global PV module production in 2012 was 35.9 GWp (GTM Research,
2013). Future growth scenarios differ widely, but the average forecast is for some
260 GWp of cumulative installed global PV capacity by 2020, with demand shift-
ing from Europe to countries such as China, India and the USA. These projections
show there are huge opportunities for PV in the future.
A number of factors have driven this growth. Many countries, including most
in the OECD and several developing countries, have introduced tax or regulatory
policies that favour the renewables. Modest renewable set-asides (requirements
on major utilities to source some power from renewables) are in place in many
countries with liberalised electricity supply industries, and the right to supply power
to the grid has been extended to independent power producers (IPPs), sometimes
with incentives to source electricity from renewable sources. IPPs can site their
plant close to the consumer and avoid the costs of distribution, often as significant
as the costs of generation.
The 100,000 roofs programme in Germany, which ran from 1999 to 2004, and
other national roofs programmes stimulated demand for distributed PV systems.
Feed-in tariffs (FiTs), also pioneered by the German government, have stimulated
PV installations by guaranteeing an above-market price for a fixed term for PV-
generated power to be fed into the grid. These tariffs, which are now offered in
many countries, give assurance of return to investors. Later FiT rounds generally
provide less generous rates of return, which has helped to drive system prices down,
although abrupt tariff changes can produce boom/bust cycles and create investor
uncertainty.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch01 page 3
4 M. D. Archer
Figure 1.1 Highest confirmed efficiencies for (a) ≥1 cm2 area cells (b) ≥800 cm2 area
modules (200–400 cm2 for OPV); (c) concentrator cells and modules. Source: Green et al.
(2013).
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch01 page 6
6 M. D. Archer
2 The 15 states that made up the European Community at the time of the Kyoto negotiations.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch01 page 7
electrodes coated with silver chloride or silver bromide and immersed in aque-
ous solution.3 The observation by Smith (1873) of photoconductivity in solid
selenium led to the discovery of the photovoltaic effect in a purely solid-state
device by Adams and Day (1877), who observed photovoltages in a selenium rod
to which platinum contacts had been sealed, which they (incorrectly) ascribed
to light-induced recrystallisation of the selenium. The first practical photovoltaic
device — a light meter consisting of a thin layer of selenium sandwiched between
an iron base plate and a semi-transparent gold top layer made by Fritts (1883) —
was promoted by the German industrialist Werner von Siemens as demonstrating
‘for the first time, the direct conversion of the energy of light into electrical energy’
(Siemens, 1885). Photometers based on selenium photocells were commercialised
in Germany in the 1930s, and amorphous selenium photodetectors remain of inter-
est today for medical imaging applications.
The selenium photocell is an example of a barrier layer cell, so called because
it contains an electrical barrier that is highly resistive to current flow in one direc-
tion — a rectifying junction, in modern parlance. Two further barrier layer cells,
the thallous sulphide cell (Case, 1920) and the copper oxide cell (Kennard and
Dieterich, 1917; Grondahl and Geiger, 1927), were developed during the 1920s.
Garrison (1923) and later Grondahl (1933) reviewed work on Cu2O, Garrison being
the first to report the logarithmic dependence of open-circuit voltage on irradiance.
Fink and Fogle (1934) discovered the efficacy of antireflective coatings accidentally
when they found that beeswax, applied to protect cells during chemical etching,
enhanced the photocurrent. Schottky (1930) noted and explained the difference in
spectral response of ‘Vorderwandzelle’ (front-wall-illuminated cells) and ‘Hinter-
wandzelle’ (back-wall-illuminated cells) while experimenting with semitranspar-
ent metal contacts for the Cu2 O/Cu barrier cell.
Lange (1930), like Siemens, foresaw the possible application of the photo-
voltaic effect to energy conversion, pointing out that ‘using a more appropriate
semiconductor for the intermediate layer [of a barrier cell], it will be possible to
select the cell effectiveness for a specific spectral region … Use of a more appropri-
ate unipolar [contact] layer is expected to give a further increase in the efficiency of
the cell. It is then possible that efficiencies can be reached which allow direct con-
version of light into electrical energy.’ Sheldon and Geiger (1922), Geiger (1923)
and Bergmann and Hänsler (1936) observed and recorded photovoltaic effects with
a wide range of other minerals and compounds, Bergman reporting what may be
the earliest observation of semiconducting effects in organic dyestuffs. Reviews
3 Becquerel’s observation was strictly speaking a photoelectrochemical effect, but its basis — the rectify-
ing junction formed between two dissimilar electric conductors — is the same as that of the photovoltaic
effect in purely solid-state devices.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch01 page 8
8 M. D. Archer
by Winther (1928) and Crossley et al. (1967) and the book by Lange (1938) give
an account of these early devices.
The electrical barrier of barrier layer cells was originally thought to lodge
in an interfacial foreign layer of high resistivity such as an oxide, but Schottky
(1938), and independently Davydov (1939) and Mott (1939), showed that a third
phase was not necessarily involved. Rather, metal/semiconductor junctions could
in themselves be rectifying by virtue of the space-charge layer created in the semi-
conductor by electronic equilibration/charge redistribution when contact was made
with a metal of different work function. The first diffusion theory of p-n junction
rectification, which became the basis for Shockley’s theory, was published by
Davydov (1938).
Metal/semiconductor devices make inefficient solar converters because their
dark currents are relatively large and this diminishes the photovoltaic response.
Semiconductor/ semiconductor junctions are better in this regard. The first p-n
junctions to be reported were the germanium homojunctions of Lark-Horovitz’s
group at Purdue University (Benzer, 1946, 1947) and the lead sulphide quasi-
homojunction formed by pressing together two PbS wafers, one enriched with
lead and the other with sulphur (Sosnowski et al., 1947).
later by the MINP (Metal-Insulator-NP junction) cell of Green et al. (1985). From
then on, Si cell technology progressed from MINP to PESC (the Passivated Emitter
Solar Cell), to rear contact cells, to PERC (Passivated Emitter and Rear Contacts)
cells, and ultimately to the PERL (Passivated Emitter, Rear Locally-diffused) cell,
which is 24.8% efficient in monocrystalline form (Zhao et al., 1998). Green (1999)
describes the development of silicon solar cell technology in more detail.
A small terrestrial market for c-Si modules started to appear for remote, small-
power applications and demonstration projects in the 1970s, stimulated by the oil
price hikes of that time (Green, 1993; Treble, 1998). Successive markets opened
up for silicon PV as costs reduced. In 1970, c-Si cells for use in space cost several
hundred dollars per watt. By the mid-1970s, the efforts of Elliot Berman and his
Solar Power Corporation (backed by Exxon) had reduced the cost of cells made
specifically for terrestrial applications to $20 Wp−1 . In round terms, the RAPS
(remote area power supplies) market opened up in the 1980s at module costs of
$10 Wp−1. Solar lighting in grid-remote locations opened up in the early 1990s
at $5 Wp−1 . Multicrystalline silicon (mc-Si) cells entered the PV market in 1981
and captured an increasing market share as module efficiencies improved from an
early 10–12% to ∼16–18% today.
Until the mid-2000s, sufficient high-quality material for c-Si cell manufacture
was available from the electronics industry as cut-price off-spec and waste polysil-
icon (small-grained, polycrystalline silicon of 9N purity).4 This was mostly made
from metallurgical-grade silicon (MG-Si) by the energy-intensive Siemens pro-
cess, which dates back to the 1940s. The main method for producing single-crystal
Si from polysilicon is the even more venerable Czochralski (CZ) method, in which
a rod of monocrystalline Si is grown by slowly withdrawing a seed crystal from a
polysilicon melt. The CZ method dates from the 1916 work of the Polish scientist
Jan Czochralski, co-founder of the German Metallurgical Society, and is also an
energy-intensive process. Cutting the CZ rod into thin wafers to make c-Si cells
incurs further losses and wastage of material, and fabricating cells from wafers is
a complex and multi-step process. Multicrystalline silicon (mc-Si) is commonly
made by the Bridgman method of directional crystallisation: silicon melt is poured
into a cast and solidified from the bottom up by extracting heat from the crucible
base, yielding blocks with columnar crystal growth that are then cross-sectioned
into large-grained wafers.
By 2005, the market for PV had grown to the point where Si PV manufac-
turers were competing against chip companies for electronic-grade material. This
4 The purity of silicon is often expressed as the total number N of nines in 99.99 … %.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch01 page 10
10 M. D. Archer
at IBM’s Thomas Watson Research Center and Alferov et al. (1967) at the Ioffe
Institute of Leningrad (as it then was).5 This was quickly followed by the p-
AlGaAs/ p-GaAs/n-GaAs heteroface cell developed by Woodall and Hovel (1972)
at the Thomas Watson Research Center, which in 1977 achieved a record single-
junction efficiency for its time of 21.9%.
Tandem solar cells were proposed as far back as 1955 by Jackson (1955) and
later by Wolf (1960). However, efficient III–V tandem cells were not achieved
until the late 1980s because of difficulties in making high-performance stable tun-
nel junctions and the effects of oxygen-related defects in the AlGaAs window layer
(Ando et al., 1987). Yamaguchi et al. (1987) proposed the use of double-hetero
(DH) structure tunnel diodes as optically and electrically low-loss cell intercon-
nects. Olson et al. (1990) achieved an efficiency of 27.3% in a cell with a Ga0.5In0.5 P
homojunction grown epitaxially on a GaAs homojunction with a GaAs tunnel junc-
tion interconnect, and Japan Energy (Takamoto et al., 1997) broke the (1 Sun)
30% barrier with their InGaP/GaAs tandem cell, which had a DH tunnel junction
interconnect in which the InGaP layers were surrounded by high-bandgap AlInP
barriers.
The main drawback of III–V cells is their high cost. Aluminium is so reactive
in the vapour phase that it is not possible to prepare AlGaAs layers by conventional
chemical vapour deposition using elemental sources. Rather it is necessary to use
the slower and more expensive method of metal-organic chemical vapour epitaxy
(MOVPE), using sources such as trimethyl aluminium. Cell stacks are usually
grown epitaxially on substrates such as germanium. However, the high efficiency,
good radiation hardness of InP and InGaP (Yamaguchi et al., 1997) and low series
resistance of III–V MJ cells have allowed them to dominate two niche markets:
provision of PV power in space and use in concentrator PV systems (CPV). III–V
MJ concentrator cells have the highest conversion efficiencies in their class of any
PV technology. Boeing Spectrolab broke the 40% efficiency barrier in 2006 with
their 3-junction GaInP/GaInAs/Ge concentrator cell, which was 40.7% efficient
under 240 Suns (King et al., 2007). At the time of writing, the world record holder
for CPV is the Fraunhofer Institute’s 4-junction cell, which is 44.7% efficient
under 297 Suns (Fraunhofer, 2013). In theory, efficiencies of 50% or more are
possible with MJ concentrator cells with four or more junctions (Yamaguchi et al.,
2008).
5 Zhores Alferov won the 2000 Nobel Prize for Physics for his development of heterostructures and is
probably the only solid-state physicist to have an asteroid named after him.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch01 page 12
12 M. D. Archer
In the early 2000s, the combination of the relatively high cost of flat-plate c-Si
PV systems and the impressive improvements in III–V MJ cell efficiency stimu-
lated a number of other commercial CPV ventures. By mid-2011 the cumulative
global installed CPV capacity was 23 MWp, Spain accounting for 70% of this
total.
2012 has seen a more mixed picture. Some commentators predict contin-
ued market growth: Lux Research (2012) saw the market growing to 697 MWp
p.a. by 2017 and IMS Research (2012) predicted that cumulative installations
would reach 1.2 GWp by 2016. However, 2012 has proved to be as difficult a
market for CPV as it has for conventional PV, and commercial CPV ventures have
retrenched. Amonix, which announced a world record CPV module efficiency of
33.5% in October 2012, closed its manufacturing facility in North Las Vegas in
July. The development of III–V multijunction cells of even higher performance
may tilt the economics back in favour of CPV: King et al. (2012) calculate that
50% efficient MJ CPV cells would be the lowest cost option for solar electric-
ity generation in high direct normal irradiance regions. Other commentators are
forecasting that CPV will be unable to compete with the continued low price of
flat-plate PV.
14 M. D. Archer
fell from favour. CdS lives on, however, as the window layer of CdTe and CIGS
cells.6
6 Cadmium sulphide also lives on in the paintings of impressionists such as Monet, whose favourite
yellow pigment it was.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch01 page 15
active top layer in the face of the material’s high surface recombination velocities.
The way forward proved to be the n-CdS/ p-CdTe heterojunction cell, in which
CdTe forms the active, light-absorbing base layer and CdS the front window layer.
The n-CdS/ p-CdTe device structure combines good optical transparency with suf-
ficiently close lattice and thermal matching to form a ‘good’ (spike-free) junction,
albeit after a special activation process. Single-crystal n-CdS/ p-CdTe cells of up to
8% efficiency were made in the 1970s (Saraie et al., 1972; Yamaguchi et al., 1977;
Mitchell et al., 1977), and this good performance allied to ease of junction forma-
tion and tolerance to materials purity that caught the attention of industry. General
Electric was an early market leader in CdTe cell manufacture (Cusano, 1962, 1966).
BP Solar followed in the 1990s with a research programme at Sunbury-on-Thames
near London on electrodeposition of CdTe cells, and later module production in
the US, but axed this programme in late 2002. Dieter Bonnet, who co-authored the
account of CdTe PV in Chapter 5, was one of the pioneers of this field. With his
co-worker Rabenhorst, he was the first to report the all-thin-film CdS/CdTe hetero-
junction cell (Bonnet and Rabenhorst, 1972) and he founded ANTEC in Kelkheim,
Germany, which for some years fabricated 7%-efficient cells deposited by closed
space sublimation (CSS).
The CdTe market phenomenon of the late 2000s — First Solar of Tempe,
Arizona — was founded by entrepreneur Harold McMaster in 1990 and is now
owned by the investment arm of the family that owns Walmart. The company has
focussed aggressively on cost reduction, with impressive results. In 2009, First
Solar’s CdTe modules were the first to break the US$1 Wp−1 cost barrier (one PV
equivalent of the 4-minute mile) and the company was the top PV supplier of that
year. Since then, the falling price of c-Si modules has eroded the value proposition
for CdTe, and First Solar, like other PV manufacturers, has had to retrench.
Best research-cell efficiency plateaued at 16.7% for a decade but has now
reached 18.7% in a cell made with First Solar’s commercial-scale manufacturing
equipment and materials (First Solar, 2013), and module efficiencies of above
11% are routinely achieved (Semiconductor Today, 2011). Given that the CdTe
bandgap of 1.44 eV is almost ideally matched to the terrestrial solar spectrum, cell
efficiencies of >20% should be possible.
Cadmium is highly toxic, and tellurium is toxic if ingested. However, fears
about the safety of CdTe technology have been largely allayed by tests showing that
module encapsulation in glass prevents escape of CdTe to the environment, though
the safe disposal of panels will require attention. Te is currently classified as an
extremely rare element (1–5 ppb in the Earth’s crust) and this could cause supply
difficulties if CdTe PV were to become a mainstream power provider. However, it
has recently been discovered that some undersea ridges are rich in the element.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch01 page 16
16 M. D. Archer
18 M. D. Archer
electrons having diffusion lengths of ∼100 nm. Moreover, little energy is lost in
separating the charge carriers (as compared with the losses incurred in exciton dis-
sociation in OPV) so high open-circuit voltages are achieved. Finally, the materials
can be prepared in tailored compositions and absorption thresholds by low-cost,
low-temperature solution methods.
The first cells that were made were not stable because the perovskites were
attacked by the liquid electrolyte of the DSSC . However, it was soon discovered
that this could be substituted by a solid-state hole conductor, and then that this
could be omitted as the materials are ambipolar and carry both holes and electrons
(Etgar et al., 2012). Other features of the DSSC also proved expendable: Henry
Snaith’s group at Oxford showed that by using CH3 NH3 PbI2 Cl the mesostructured
TiO2 scaffold could be replaced by alumina (Lee et al., 2012), possibly because
of the order of magnitude increase in the exciton/charge carrier diffusion length in
CH3 NH3 PbI2 Cl compared with CH3 NH3 PbI3 (Stranks et al., 2013). It was also
found that the perovskite itself did not have to be nanostructured: a simple planar
heterojunction cell consisting of a few hundred nanometres of vapour-deposited
perovskite sandwiched between two charge-selective electrodes can have efficien-
cies of over 15% (Liu et al., 2013). Michael Grätzel’s group (Burshka et al., 2013),
working with a mesostructured TiO2 scaffold, showed that sequential deposition
of the perovskite, first laying down PbI2 and then converting it to perovskite by
exposure to CH3 NH3 I in solution, yields more consistently high-performance cells
than co-deposition.
Perovskite cells will certainly be an active area of research in the next few
years, with some commentators predicting that efficiencies of over 20% may be
possible (Park, 2013). However, many questions about stability, durability and
device repeatability remain to be addressed.
20 M. D. Archer
‘built-in’ electric field. When the cell absorbs light, excitons (electron-hole pairs)
are created. In broadband (most inorganic) semiconductors, these excitons very
quickly dissociate into mobile electrons and holes, which flow in opposite direc-
tions across the junction. In narrow-band and organic semiconductors, excitons do
not dissociate spontaneously but are separated at the junction. In either case the flow
of absorbed photons is converted into a flow of DC power from the illuminated cell.
The crystalline silicon (c-Si) cell has a simple junction structure, and provides
a good model with which to explore the PV effect. Figure 1.2 shows the essential
features of these cells, which are typically square wafers of dimensions ∼15 cm ×
15 cm × 0.2 mm. The top (emitter) region is a 0.2–0.3 µm thick layer of n-type
silicon, and the base region is a 300 µm thick layer of p-type silicon.7 The work
function of the p material is greater than that of the n material, so the two layers
reach electronic equilibrium (in the dark) by the transfer of some electrons from
the n to the p side. The structure as a whole remains electrically neutral, but the
junction region contains an electric double layer, consisting of two space-charge
regions or depletion regions (DRs), as shown in Fig. 1.3. These DRs are typically
less than a micron thick, and the charges they contain are those of the ionised
dopants (P+ and B− in the case of c-Si). Beyond the base-layer DR in the c-Si cell
(and some other cells) lies a quasineutral region (QNR) — a region that contains
no space charge.
Figure 1.4 shows what happens in the illuminated c-Si cell. The absorption
of photons of energy greater than the bandgap energy of silicon promotes elec-
trons from the valence band to the conduction band, creating hole-electron pairs
7 c-Si cells are usually configured n-on- p because this best suits the properties of silicon, but some other
p-n cells are configured p-on-n. These cells are also quite thick, because c-Si absorbs light relatively
weakly. Most other cells are much thinner.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch01 page 21
Figure 1.3 Cross section through a p-n homojunction cell, showing the electrical double
layer consisting of ionised dopant atoms (denoted + and –) in the junction region, the two
depletions regions (DRs) that contain equal and opposite quantities of junction charge, and
the base-layer quasineutral region (QNR).
Figure 1.4 Generation and movement of free carriers in a p-n junction cell.
throughout the illuminated part of the cell, which in c-Si cells extends well into
the base layer. In c-Si and most other semiconductors, these hole-electron pairs
quickly dissociate into ‘free’ carriers — mobile holes and electrons that move
independently of each other. Those free carriers that approach the junction come
under the influence of the built-in electric field, which sweeps electrons from the
p to the n side, and holes from the n to the p side.
22 M. D. Archer
(a) (b)
Figure 1.5 Energy band structure of a p-n homojunction in the dark: (a) in uncharged
blocks of p-type and n-type semiconductors before contact, showing the conduction and
valence-band energies E c and E v , the forbidden gap E g and the Fermi levels E Fn and E Fn
(red dashed lines) in the n and p phases; (b) across the p-n homojunction after contact and
equilibration of the two phases, showing the electric double layer formed by transient charge
transfer, the depletion regions (DRs) and quasi-neutral regions (QNRs) and the common
Fermi level E F throughout the device.
(Fig. 1.5a) have the same conduction and valence band-edge energies E c and E v ,
separated by the forbidden gap E g , but different work functions p and n , and
therefore different Fermi levels E F p and E Fn .8 In the equilibrated cell (Fig. 1.5b),
the Fermi level E F is the same throughout the device but the band-edge energies E v
and E c (in common with all the energy levels of the semiconductor) bend across
the junction in response to the local electric field. Inspection of Fig. 1.5 shows that
the equilibrium band-bending energy is qVob is related to the difference in the work
functions of the (separate, uncharged) materials by
q Vbo = n − p (1.1)
Since the Fermi level in a doped semiconductor normally lies within the forbid-
den gap but near the majority-carrier band edge, qV ob is normally slightly smaller
than the bandgap energy E g . Figure 1.6 shows how the band bending is affected
and a current is caused to flow when a bias voltage V j is applied across the cell
in the dark. At equilibrium (Fig. 1.6a), no net current9 flows through any part of
8 The Fermi level is the energy for which the probability of a state being occupied by an electron is
exactly one-half. In an intrinsic (undoped) semiconductor, the Fermi level falls in the middle of the
forbidden gap. In a lightly doped semiconductor, the Fermi level remains within the forbidden gap but
is near the majority-carrier band edge. In a heavily doped semiconductor, the Fermi level lies within
the majority-carrier band.
9 All the currents given the symbol i in Figs. 1.6–1.8 are strictly speaking current densities.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch01 page 23
the cell. However, small, balanced fluxes of electrons in the conduction band and
holes in the valence band pass each way across the junction. These are referred to
as generation and recombination currents. The (thermal) generation currents i h,gen
and i e,gen shown in Fig. 1.6a come from the minority carriers (electrons in the p
side and holes in the n side) generated throughout the device, albeit at a minuscule
rate, by thermal excitation. Those minority carriers that reach the junction with-
out recombining are swept across it in opposite directions by the strong electric
o
field. The recombination currents i h,rec o
and i e,rec also shown in Fig. 1.6a come
from majority carriers (holes in the p side and electrons in the n side) that flow
‘up’ the band-bending barrier (this is energetically unfavourable, but entropically
favourable because the carriers move from a region of high to low concentration).
At equilibrium, the generation and recombination currents in each band
exactly balance each other. The sum of the hole and electron thermal generation
currents is called the saturation current density i o of the junction.
When a forward bias10 voltage V j is applied across the junction of the dark
cell, the barrier height is reduced to qVb = q(Vbo − V j ), as shown in Fig. 1.6b. This
does not affect the generation currents, but it strongly increases the recombination
currents. The net current across the junction, which is the difference between the
recombination current and the generation current, is called the dark current or
10 Forward biasing a junction means applying a voltage across the device that lowers the band-bending
barrier. Reverse biasing means applying a voltage in the opposite direction.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch01 page 24
24 M. D. Archer
junction current i j .
i j (V j < 0) = −i o (1.4)
where β is called the diode ideality factor. For an ideal junction, in which no
injected carriers recombine in the junction, β = 1. For a non-ideal junction, in
which some carriers do recombine in the junction, 1 < β < 2. For many cells,
Eq. (1.5) is better written as the two-diode equation
where the first term corresponds to carriers that move across the junction without
recombining, and the second to the carriers that recombine in mid-gap. Regardless
of the exact form of the diode equation, all PV cells behave as rectifiers in the dark,
showing highly non-linear current–voltage characteristics similar to that labelled
‘dark’ in Fig. 1.8. Junctions must show rectifying properties in the dark if they are
to show photovoltaic properties in the light.
(a) (b)
Figure 1.7 Illuminated p-n homojunction cell (a) at open circuit, showing the photocurrents
as red dotted lines and the hole and electron quasi-Fermi levels E Fh and E Fe across the junction
region as red dashed lines; (b) at short circuit, assuming that the cell has no internal resistance
(i.e. that carrier mobility is infinite).
from the QNRs towards the junction, where they are swept across it by the strong
junction field. These fluxes of photogenerated minority carriers give rise to the
photogeneration currents i e, ph and i h, ph shown in Fig. 1.7a, consisting respec-
tively of photogenerated electrons drifting from the p to the n side of the junction
and photogenerated holes drifting the other way. The sum of the two is the overall
photocurrent i ph .
i ph = i h, ph + i e, ph (1.7)
The photocurrent is directly proportional to the absorbed photon flux but inde-
pendent of bias (provided that the junction field is always high enough to sweep
carriers across the junction). At open circuit (Fig. 1.7a), no current is drawn from the
cell and the photocurrent must be balanced by the recombination current. The junc-
tion self-biases in the forward direction by the open-circuit voltage Voc , at which
point the recombination (junction) current exactly opposes the photocurrent, i.e.
i ph − i j (Voc ) = 0 (1.8)
As shown in Fig. 1.7a, the hole and electron quasi-Fermi levels E Fh and E Fn
diverge across the junction and converge past the quasi-neutral regions, and qVoc is
the difference between the Fermi levels on the two far sides of the junction. Since
metal contacts always equilibrate with the local majority carrier Fermi level, Voc
is an observable output voltage.
Figure 1.7b shows what happens when the illuminated cell is short-circuited.
The cell delivers maximum current but at zero output voltage. Provided internal
resistance effects are negligible, the junction bias V j is also zero, so the band
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch01 page 26
26 M. D. Archer
i sc = |i ph | − i o (1.9)
Under normal operating conditions, the band bending and junction current are
intermediate between the open-circuit and short-circuit cases, and the cell delivers
current i at output voltage V ≈ V j , where i is given by
i = i ph − i j (V j ) (1.10)
Figure 1.8 Current–voltage curves in the dark and the light for a cell that shows super-
position (i.e. one in which the photogenerated current is bias-independent), showing the
short-circuit and maximum-power currents i sc and i mp , the open-circuit and maximum-
power voltages Voc and Vmp , and the maximum power point ().
11 If the cell has significant internal resistance, the output voltage V drops below the junction voltage
V j , and a small forward bias remains across the junction when the cell is short-circuited.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch01 page 27
that of the majority carriers, do not show superposition because the majority-carrier
concentrations and fluxes are not then the same in the light and the dark. Cells with
significant internal series resistance or shunt conductance also depart from super-
position.
The output power is the product iV, which is the area of a rectangle of sides i
and V inscribed in the i –V curve. The power is zero for both the open-circuit and
short-circuit conditions. The maximum-power condition is reached where the area
i mp Vmp (shaded in Fig. 1.8) is a maximum. The fill factor ηfill is a measure of the
squareness of the i –V curve and is defined as
i mp Vmp
ηfill = (1.12)
i sc Voc
In efficient cells, the fill factor is around 0.7–0.8. In poor cells, it can be 0.5 or
lower.
By setting i = 0, V = Voc in Eq. (1.11) and rearranging, the open-circuit
voltage of the illuminated cell is found as
βkT i ph βkT i ph
Voc = ln 1 + ≈ ln (1.13)
q io q io
For good performance, i ph and Voc must be as large as possible. The max-
imum value of i ph would be obtained if all photogenerated electron-hole pairs
were collected as photocurrent, and i ph can achieve over 90% of this limit if light
absorption and minority carrier collection are both highly efficient. If there are no
sources of voltage in the cell other than at the junction, the limiting value of Voc is
the built-in voltage Vbo , corresponding to complete flattening of the bands across
the junction.12 This would only happen under extremely intense illumination, and
1 Sun Voc values are usually no more than ∼0.75 Vbo ; for the best GaAs cells,
Voc ≈ 0.79E g . For a high open-circuit voltage, Vbo should be as large as possible
12 In lightly doped Si cells, V can exceed V o at the junction, with the bulk going into high injection
oc b
and an additional voltage across the rear back-surface field: SunPower cells effectively work in this
way.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch01 page 28
28 M. D. Archer
given the semiconductor bandgap, so the work function difference between the two
sides of the junction should be as large as possible.
Inspection of Eq. (1.13) shows that Voc increases as the saturation current i o
decreases. Interestingly, i o has no absolute minimum value. In thin cells with well-
passivated surfaces, i o can be driven down toward zero, and Voc towards its upper
limit of Vbo . In thicker cells in which volume recombination occurs, the lower limit
on i o is determined by the rate of radiative recombination of minority carriers.
Usually nonradiative recombination also occurs and this raises i o by several orders
of magnitude, lowering Voc accordingly.
13 An ideal isotropic cell is one in which electrons and holes are thermalised to the band edge, the only
decay channel for excited states is radiative recombination, and light can enter the cell at all forward
angles.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch01 page 30
30 M. D. Archer
Figure 1.9 Common PV junction types. (a) p-n homojunction, formed within a single
semiconductor of bandgap E g ; (b) p-i-n junction, formed within a single semiconductor of
bandgap E g ; (c) anisotype P-n heterojunction formed between semiconductors of bandgaps
E g1 and E g2 , showing a valence-band spike E v and a conduction-band notch E c ; (d)
P + - p-n heteroface junction; (e) MS junction between a metal M and an n-type semicon-
ductor S; (f) MIS junction with a thin layer of an insulator I interposed between M and S; (g)
organic cell containing an organic co-polymer blend between a top transparent conducting
oxide (TCO) electrode and a metal electrode M. All these junctions are shown at equilibrium
in the dark, so the Fermi levels, shown by the red dashed lines, are the same throughout
each junction.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch01 page 31
n-layers of the same semiconductor. The i -layer behaves like the dielectric in
a capacitor, effectively stretching the electric field of an ordinary p-n junction
across itself. The extended electric field throughout the i -layer, which is where
most of the light is absorbed, aids in the collection of photogenerated carriers by
adding a component of drift (migration in an electric field) to their normal diffusive
motion. This is the junction type in amorphous silicon (a-Si:H) cells, and it was
adopted because the high density of states and traps in this disordered material
causes carrier mobilities to be low.
Heterojunctions (Fig. 1.9c) are junctions formed between two chemically dif-
ferent semiconductors with different bandgaps. The larger bandgap material is
often denoted by writing its conductivity type as upper-case N or P, and the
smaller bandgap material by a lower-case letter. In the frontwall configuration,
the top layer is the main light-absorbing layer, and light enters through its front
surface. In the backwall configuration, the top layer is an optical window material
of wide bandgap through which light passes to enter the light-absorbing base layer
at the junction. Heterojunctions may be anisotype, meaning that the two semicon-
ductors have opposite conductivity types, or isotype, meaning that they have the
same conductivity type.
The advantage of this junction type is that it allows semiconductors which have
good light absorption and carrier lifetime properties, but can only be doped n- or
p-type, to be used in solar cells. Its disadvantages are first, that any significant lat-
tice mismatch between the two materials creates numerous junction defects, which
diminishes the photovoltage, and second, that the energy band mismatch between
the two materials creates notches or spikes in the junction band-edge profile. Fig-
ure 1.9c shows a poor junction with a pronounced valence-band spike, which would
seriously impede the collection of photogenerated holes from the n-semiconductor.
The two most successful heterojunctions for PV applications, n-CdS/ p-CdTe and
n-CdS/ p-CuInSe2 , are both N- p anisotype backwall devices with good lattice
matches between the two materials, and favourable band alignment with no spike
presented to carriers arriving at the junction. Heteroface junctions or buried homo-
junctions (Fig. 1.9d) have a p-n homojunction fronted by a highly conducting
window layer (face) semiconductor of larger bandgap (E g1 > E g2 ). This window
layer, although not photovoltaically active, is beneficial. As well as acting as current
collector from the top layer, it passivates its surface and minimises front-surface
recombination. The p+ -Alx Ga1−x As/ p-GaAs/n-GaAs cell is an important exam-
ple of this cell type. An unprotected p-n GaAs homojunction would not make an
efficient solar cell because the top layer of this direct-gap material must be very
thin to allow light to penetrate to the junction region; the dark current would then
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch01 page 32
32 M. D. Archer
14 Rather than in accumulation, which would create a photovoltaically inactive ohmic contact.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch01 page 33
materials are sufficiently soft that the cell constituents may be co-blended to create
a bulk heterojunction cell with interfaces throughout the device. This increases
the chance of exciton formation within a diffusion length of the nearest junction.
However, even the best OPV cells have low open-circuit voltages for their effec-
tive bandgap, arising from their poor radiative efficiencies (i.e. high nonradiative
recombination rates), which is due to their blended structure.
References
Adams W. G. and Day R. E. (1877), ‘The action of light on selenium’, Proc. Roy. Soc. A25,
113–117.
Ando K., Amano C., Sugiura H., Yamaguchi M. and Salates A. (1987), ‘Nonradiative e-h
recombination characteristics of mid-gap electron trap in Al x Ga1−x As (x = 0.4)
grown by molecular beam epitaxy’, Jpn. J. Appl. Phys. 26, L266–L269.
Archer M. D., Bolton J. R. and Siklos S. T. C. (1996), A review of analytic solutions for
a model p-n junction under low-injection conditions,Solar Energy Mater. Solar Cells
40, 133–176.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch01 page 34
34 M. D. Archer
Bailat J., Fesquet L., Orhan J.-B., Djeridane Y., Wolf B., Madhiger P., Steinhauser J., Benagli
S., Borrello D., Castens L., Monteduro G., Marmelo M., Dehbozorgi B., Vallat-Sauvain
E., Multone X., Romang D., Boucher J.-F., Meier J., Kroll U., Despeisse M., Bugnon
G., Ballif C., Marjanovic S., Kohnke G., Borrelli N., Koch K., Liu J., Modavis R.,
Thelen D., Vallon S., Zakharian A. and Wiedman D. (2010), ‘Recent developments
of high-efficiency micromorph tandem solar cells in KAI-M PECVD reactors’, Proc.
5th. World Conf. on Photovoltaic Energy Conversion, Valencia, Spain, 6–10 September
2010.
Barnham K. W. J. and Duggan G. (1990), ‘A new approach to high-efficiency multi-bandgap
solar cells’, J. Appl. Phys. 67, 3490–3493.
Becquerel A. E. (1839), ‘Recherches sur les effets de la radiation chimique de la lumière
solaire, au moyen des courants électriques’, Compt. Rend. Acad. Sci. 9, 145–149,
561–567.
Benzer S. (1946), The photo-diode and photo-peak characteristics in germanium’, Phys.
Rev. 70, 105.
Benzer S. (1947), ‘Excess-defect germanium contacts’, Phys. Rev. 72, 1267–1268.
Bergmann L. and Hansler J. ‘Lichtelektrische Untersuchungen an Halbleitern’, Zeits. Physik
100, 50– 79.
Bonnet D. and Rabenhorst H. (1972), ‘New results on the development of thin film p-
CdTe/n-CdS heterojunction solar cell’, Proc. 9th IEEE PVSC, Silver Springs, MD
(IEEE, 1972), pp. 129–132.
Brittlebank W. (2012), ‘Solar PV approaching grid parity’, www.climateaction.org, 10
August 2012, retrieved 12 January 2013.
Burschka J., Kessler D. A., Baranoff E., Cevey-Ha N. L., Yi C., Nazeeruddin M. K. and
Grätzel M. (2011), ‘Tris(2-(1H-pyrazol-1-yl)pyridine)cobalt(III) as p-type dopant for
organic semiconductors and its application in highly efficient solid-state dye-sensitized
solar cells’, J. Amer. Chem. Soc.133, 18042–18045.
Burschka J., Pellet N., Moon S.-J., Humphry-Baker R., Gao P., Nazeeruddin M. K. and
Grätzel M. (2013), ‘Sequential deposition as a route to high performance perovskite-
sensitized solar cells’, Nature 499, 316–319.
Carlson A. (1956), Research in Semiconductor Films, WADC Technical Report, Clevite
Corporation.
Carlson D. E. and Wronski C. R. (1976), ‘Amorphous silicon solar cells’, Appl. Phys. Lett.
28, 671–673.
Case T. W. (1920), “‘Thalofide cell” — a new photoelectric substance’, Phys. Rev. 15,
289–292.
Chapin D. M., Fuller C. S. and Pearson G. O. (1954), ‘A new silicon p–n junction
photocell for converting solar radiation into electrical power’, J. Appl. Phys. 25,
676–677.
China Briefing (2012), ‘China releases twelfth five-year plan on solar power development’,
www.china-briefing.com/news, 19 September 2012, retrieved 7 January 2013.
Crossley P. A., Noel G. T. and Wolf M. (1967), ‘Review and evaluation of past
solar cell development efforts’, Report prepared under contract for NASA.
http://ntrs.nasa.gov/archive/nasa/casi.ntrs.nasa.gov/19670022851_1967022851.pdf.
Cusano D. A. (1962), ‘Polycrystalline thin-film CdTe solar cells’, IRE Trans. Electron
Devices ED–9, 504.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch01 page 35
Cusano D. A. (1966), ‘The performance of thin film solar cells employing photovoltaic
Cu2−x Te–CdTe heterojunctions’, Rev. Phys. Appl. 1, 195–200.
Davydov B. I. (1938), ‘Theory of rectification in the semiconductors’, Bull. Acad. Sci.
U.S.S.R. 5–6, 625–629.
Davydov B. I. (1939), ‘Contact resistance of semiconductors’, Zh. Eksp. Teor. Fiz. 9,
451–458.
DECC (2013), UK Solar PV Strategy Part 1: Roadmap to a Brighter Future, Department
of Energy and Climate Change, October 2013.
Eight19 (2012), ‘Eight19 spins out Indigo pay-as-you-go solar’, www.eight19.com, 21
August 2012, retrieved 12 January 2013.
EPIA (2013), Global Market Outlook for Photovoltaics 2013–2017.
Etgar L., Gao P., Xue Z., Peng Q., Chandiran A. K., Liu B., Nazeeruddin M. K. and Grätzel M.
(2012), ‘Mesoscopic CH3 NH3 PbI3 /TiO2 heterojunction solar cells’, J. Amer. Chem.
Soc. 134, 17396–17399.
Fink C. G. and Fogle M. E. (1934),‘A study of cuprous oxide solid photoelectric cells’,
J. Electrochem. Soc. 66, 271–322.
First Solar (2013), ‘First Solar sets new world record for CdTe solar cell efficiency’,
http://investor.firstsolar.com press release 26 February 2013, retrieved 5 April 2013.
Fraunhofer (2013), ‘World record solar cell with 44.7% efficiency’, www.ise.fraunhofer.de
23 September 2013, retrieved 30 November 2013.
Fraunhofer Institute for Solar Energy Systems ISE (2012), Photovoltaics Report, December
11, 2012.
Fritts C. E. (1883), ‘On a new format of selenium cell, and some electrical discoveries made
by its use’, Am. J. Sci. 26, 465–472.
Garrison A. D. (1923), ‘The behaviour of cuprous oxide photo-voltaic cells’, J. Phys. Chem.
27, 601–622.
Geiger P. H. (1923), ‘Spectro-photoelectrical effects in argentite: the production of an elec-
tromotive force’, Phys. Rev. 22, 461– 469.
Ghosh A. K., Morel D. L., Feng T., Shaw R. F. and Rowe C. A. (1974), ‘Photovoltaic and
rectification properties of Al/Mg phthalocyanine/Ag Schottky-barrier cells’, J. Appl.
Phys. 45, 230–236.
Grätzel M. (2000), ‘Perspectives for dye-sensitized nanocrystalline solar cells’, Progr. Pho-
tovoltaics 8, 171–185.
Green M. (1993), ‘Silicon solar cells: evolution, high-efficiency design and efficiency
enhancements’, Semiconductor Sci. Technol. 8, 1–12.
Green M. A. (2000), ‘Third generation photovoltaics: advanced structures capable of high
efficiency at low cost’, 16th. European Photovoltaic Solar Energy Conf., Glasgow, 1–5
May 2000.
Green M. A. (2009), ‘The path to 25% silicon solar cell efficiency: history of silicon cell
evolution’, Progr. Photovoltaics 17, 183–189.
Green M. A., Blakers A. W. and Osterwald C. R. (1985), ‘Characterization of high-efficiency
silicon solar cells’, J. Appl. Phys. 58, 4402–4408.
Green M. A., Emery K., Hishikawa Y., Warta W. and Dunlop E. D. (2013), ‘Solar cell
efficiency tables (version 41)’, Progr. Photovoltaics 21, 1–11.
Gremmelmaier R. (1955), ‘Gallium-arsenic photoelement’, Z. Naturforsch. A 19,
501–502.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch01 page 36
36 M. D. Archer
Lee M. M., Teuscher J., Miyasaka T., Murakami T. N. and Snaith H. J. (2012), ‘Effi-
cient hybrid solar cells based on meso-superstructured organometal halide perovskites’,
Science 338, 643–647.
Lindmeyer J. and Alison J. F. (1973), ‘The violet cell: an improved silicon solar cell’,
COMSAT Technical Review 3, 1–22.
Liu M., Johnston M. B. And Snaith H. J. (2013), ‘Efficient planar heterojunction perovskite
solar cells by vapour deposition’, Nature 501, 395–398.
Lux Research (2012), Putting High-Concentrating Photovoltaics into Focus, June 2012.
Mahdi W. (2012), ‘Qatar to tender 200 megawatt of solar power projects in 2013’,
www.bloomberg.com, 3 December 2012, retrieved 3 January 2013.
Meier J., Dubail S., Flückiger R., Fischer D., Keppner H. and Shah A. (1994), ‘Intrinsic
microcrystalline silicon (mc-Si:H) — a promising new thin film solar cell material’,
Proc. 1st. World Conf. on Photovoltaic Energy Conversion, Hawaii, pp. 409–412.
Mishra A., Fischer M. K. R. and Bauerle P. (2009), ‘Metal-free organic dyes for dye-
sensitized solar cells: from structure–property relationships to design rules’, Angew.
Chem. Int. Ed. 48, 2474–2499.
MIT (2011), ‘Why BP Solar failed’, www.technologyreview.com 21 December 2011,
retrieved 26 January 2014.
Mitchell K., Fahrenbruch A. L. and Bube R. H. (1977), ‘Evaluation of the CdS/CdTe het-
erojunction solar cell’, J. Appl. Phys. 48, 4365–4371.
Mitzi D. B., Feild C. A., Harrison W. T. A. and Guloy A. M. (1994), ‘Conducting tin halides
with a layered organic-based perovskite structure’, Nature 369, 467–469.
Mott N. F. (1939), ‘Copper–cuprous oxide photocells’, Proc. Roy. Soc. A171, 281–285.
Ohl R. S. (1941), ‘Light-sensitive electric device’, U.S. Patent No. 2,402,662; ‘Electrical
translating device utilizing silicon’, U.S. Patent No. 2,402,839; ‘Light-sensitive device
including silicon’, U.S. Patent No. 2,443,542.
Olson J. M., Kurtz S. R., Kibbler A. E. and Faine P. (1990), ‘A 27.3% efficient
Ga0.5 In0.5 P/GaAs tandem solar cell’, Appl. Phys. Lett. 56, 623–625.
O’Regan B. and Grätzel M. (1991), ‘A low-cost, high-efficiency solar cell based on dye-
sensitized colloidal TiO2 films’, Nature 353, 737–740.
Park N.-G. (2013), ‘Organometal perovskite light absorbers toward a 20% efficiency low-
cost solid-state mesoscopic solar cell’, J. Phys. Chem. Lett. 4, 2423–2429.
Pochettino A. (1906), ‘Sul comportamento foto-elettrico dell’ antracene’, Acad. Lincei Ren-
diconti 15, 355–363.
PV Magazine (2013), ‘Solar stocks: China left for dead; FSLR, SPWR and WFR to rise’,
www.pv-magazine.com, 2 April 2013, retrieved 6 April 2013.
PV Tech (2012), ‘Around 90% of Chinese polysilicon producers stop production’, www.pv-
tech.org, 13 December 2012, retrieved 8 January 2013.
Reynolds D. C., Leies G., Antes L. L. and Marburger R. E. (1954), ‘Photovoltaic effect in
cadmium sulfide’, Phys. Rev. 96, 533–534.
Saraie J., Akiyama M. and Tanaka T. (1972), ‘Epitaxial growth of cadmium telluride by a
closed-space technique’, Jpn. J. Appl. Phys. 11, 1758–1759.
Schottky W. (1930), ‘Über den Entstehungsort der Photoelektronen in Kupfer–
Kupferoxydul–Photozellen’, Zeit. Tech. Phys. 11, 458–461.
Schottky W. (1938), ‘Halbleitertheorie der Sperrschicht’, Naturwiss. 26, 843.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch01 page 38
38 M. D. Archer
Semiconductor Today (2011), ‘First Solar raises CdTe PV cell efficiency record from 16.7%
to 17.3%’, www.semiconductor-today.com, 26 July 2011, retrieved 12 January 2013.
Shay J. L., Wagner S. and Kasper H. M. (1975), ‘Efficient CuInSe2 /CdS solar cells’, Appl.
Phys. Lett. 27, 89–90.
Sheldon H. H. and Geiger P. H. (1922), ‘The production of an E.M.F. on closed circuit, by
a light effect on argentite’, Phys. Rev. 19, 389– 390.
Siemens W. (1885), ‘On the electromotive action of illuminated selenium discovered by
Mr. Fritts, of New York’, Van Nostrand’s Engineering Magazine 32, 392n.
Smith W. (1873), ‘The action of light on selenium’, J. Soc. Telegraph Engineers 2, 31–33.
Söderström K., Bugnon G., Biron R., Pahud C., Meillaud F., Haug F.-J. and Ballif C. (2012),
‘Thin-film silicon triple-junction solar cell with 12.5% stable efficiency on innovative
flat light-scattering substrate’, J. Appl. Phys. 112, 114503–114504.
Solar Frontier (2013), ‘Solar Frontier achieves record 19.7% CIS efficiency’, www.solar-
frontier.com, 8 January 2013, retrieved 11 January 2013.
Sosnowski L., Starkiewicz J. and Simpson O. (1947), ‘Lead sulfide photoconductive cells’,
Nature 159, 818–819.
Spanggaard H. and Krebs F. C. (2004), ‘A brief history of the development of organic and
polymeric photovoltaics’, Solar Energy Mater. Solar Cells 83, 125–146.
Spear W. E. and Le Comber P. G. (1975), ‘Substitutional doping of amorphous silicon’,
Solid State Commun. 17, 1193–1196.
Stranks S. D., Peron G. E., Grancini G., Menelaou C., Alcocer M. J. P., Leijtens T.,
Herz L. M., Petrozza A. and Snaith H. J.(2013), ‘Electron-hole diffusion lengths
exceeding 1 micrometer in an organometal trihalide perovskite absorber’, Science 342,
341–344.
Takamoto T., Ikeda E., Kurita H. and Ohmori M. (1997), ‘Over 30% efficient InGaP/GaAs
tandem solar cells’, Appl. Phys. Lett. 70, 381–383.
Tang C. W. and Albrecht A. C. (1975a), ‘Photovoltaic effects of metal–chlorophyll-a–metal
sandwich cells’, J. Chem. Phys. 62, 2139–2149.
Tang C. W. and Albrecht A. C. (1975b), ‘Transient photovoltaic effects in metal–chlorophyll-
a–metal sandwich cells’, J. Chem. Phys. 63, 953–961.
Treble F. (1998), ‘Milestones in the development of crystalline silicon solar cells’, Renew-
able Energy 15, 473–478.
Wagner S., Shay J. L., Migliorato P. and Kasper H. M. (1974), ‘CuInSe2 /CdS heterojunction
photovoltaic detectors’, Appl. Phys. Lett. 25, 434–435.
Welker H. (1954), ‘Semiconducting intermetallic compounds’, Physica 20, 893–909.
Winther C. (1928), ‘Über den Becquereleffekt. I.’, Zeits. Physik. Chem. 131, 205– 213.
Wolf M. (1960), ‘Limitations and possibilities for improvement of photovoltaic energy
converters. Part I: considerations for Earth’s surface operation’, Proc. Inst. Radio Eng.
48, 1246–1263.
Woodall J. M. and Hovel H. J. (1977), ‘Isothermal etchback-regrowth method for high-
efficiency Ga1−x /Alx As–GaAs solar cells’, Appl. Phys. Lett. 30, 492–493.
Yamaguchi K., Nakayama N., Matsumoto H. and Ikegami S. (1977), ‘Cadmium sulphide–
cadmium telluride solar cell prepared by vapor phase epitaxy’, Jpn. J. Appl. Phys. 16,
1203–1211.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch01 page 39
Yamaguchi M., Amano C., Sugiura H. and Yamamoto A. (1987), ‘High efficiency
AlGaAs/GaAs tandem solar cells’, Proc. 19th IEEE Photovoltaic Specialists Conf.,
IEEE, New York, pp. 1484–1485.
Yamaguchi M., Okuda T., Taylor S. J., Takamoto T., Ikeda E. and Kurita H. (1997), ‘Superior
radiation-resistant properties of InGaP/GaAs tandem solar cells’, Appl. Phys. Lett. 70,
1566–1568.
Yamaguchi M., Takamoto T. and Araki K. (2008), ‘Present and future of super high efficiency
multi-junction solar cells’, Proc. SPIE 6889, Physics and Simulation of Optoelectronic
Devices XVI, 688906.
Yang G. (2012), ‘Should China bail out its solar PV industry?’, www.chinadialogue.net, 12
September 2012, retrieved 4 January 2013.
Yella A., Lee H.-W., Tsao H. N., Yi C., Chandiran A. K., Nazeeruddin M. K., Diau E. W.-G.,
Yeh C.-Y., Zakeeruddin S. M. and Grätzel M., ‘Porphyrin-sensitized solar cells with
cobalt (II/III)–based redox electrolyte exceed 12% efficiency’, Science 334, 629–634.
Yu G., Pakbaz K. and Heeger A. J. (1994), ‘Semiconducting polymer diodes: large size, low
cost photodetectors with excellent visible-ultraviolet sensitivity’, Appl. Phys. Lett. 64,
3422–3424.
Zhao J., Wang A., Green M. A. and Ferrazza F. (1998), ‘Novel 19.8% efficient “honeycomb”
textured multicrystalline and 24.4% monocrystalline silicon solar cells’, Appl. Phys.
Lett. 73, 1991–1993.
ZSW (2013), ‘ZSW produces world record thin-film solar PV cell, achieves 20.8% efficiency
and overtakes multi-crystalline silicon technology’, www.solarserver.com 24 October
2013, retrieved 30 November 2013.
May 2, 2013 14:6 BC: 8831 - Probability and Statistical Theory PST˙ws
CHAPTER 2
2.1 Introduction
Although the sunlight conversion efficiencies of most photovoltaic systems
installed to date have been quite modest, generally in the 10–15% range, there
is no fundamental reason why these cannot be very much higher. Historically, each
successful photovoltaic technology has evolved to ever-increasing energy conver-
sion efficiency, while simultaneously reducing cost. It seems reasonable to expect
this trend to continue as the industry grows, with conversion efficiency increasingly
becoming a key differentiator between technologies, as already apparent over the
last decade with the thin-film technologies. This makes it relevant to understand
fundamental constraints on efficiency and how these may be circumvented.
The publishing of the quantum mechanical theory of semiconductors (Wilson,
1931) coincided with a surge of interest in cuprous oxide solar cells (Grondahl,
1933), stimulating rapid progress in theoretical understanding of the photovoltaic
effect. The serendipitous discovery of the silicon p-n junction at Bell Laboratories
in the early 1940s due to its large photovoltages (Ohl, 1941; Riordan and Hoddeson,
1997) led to the development of p-n junction theory (Shockley, 1949) and its
extension to illuminated junctions (Cummerow, 1954a). The reporting of greatly
improved silicon cells in 1954 (Chapin et al., 1954) was followed by an estimation
of the limiting efficiency (Cummerow, 1954b) and investigations of the optimum
bandgap (Rittner, 1954; Trivich and Flinn, 1955; Loferski, 1956).
41
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch02 page 42
42 M. A. Green
100%
thermodynamic limit
74% tandem (n )
68% hot carrier
tandem (n = 6)
58% thermal, thermoPV, thermionics
54% tandem (n = 3)
49%
44% impurity PV & band, up-converters
impact ionisation
39% tandem (n = 2)
31% down-converters
single cell
Sun
0%
Figure 2.1 Limiting efficiency of various photovoltaic conversion options under global
sunlight (6000 K Sun and 300 K ambient assumed; n = number of junctions). Source: Green
and Ho (2011).
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch02 page 43
such approaches remain the single junction and the monolithic stack of multiple-
junction tandem devices that provide the main focus of this chapter and of the
present volume.
Complementing these ‘bottom-up’ limiting efficiency calculations have been
‘top-down’ thermodynamic analyses that are both device- and system-agnostic,
based purely on the energy and entropy fluxes associated with the incident sunlight
(Landsberg and Tonge, 1980). The corresponding limit for the ‘top-down’ approach
is also shown in Fig. 2.1 and lies well above the highest efficiency from the ‘bottom-
up’ approaches. This was once thought to be because the ‘top-down’ limit was
unattainable, even in principle, since the analysis failed to include unavoidable
losses (Pauwels and De Vos, 1981; Marti and Araujo, 1996). However, from the
work of Ries (1983), it becomes clear that the difference arises from the implicit
assumption of time-symmetry in the ‘bottom-up’ approaches. Photovoltaic systems
that are capable of the thermodynamically limiting performance are possible, in
principle, although the only suggested implementation, subsequently discussed, is
very complex.
Resonating with the remark by Disraeli opening this chapter, the author has
elsewhere emphasised the likely long-term evolution of photovoltaics to high-
est possible energy conversion efficiencies, as a path to lowest possible costs
(Green, 2003). Figure 2.2a effectively demonstrates the enormous leverage of high
efficiency in reducing the cost/watt of solar modules. At the system level, when
area-related installation costs are included, the leverage increases further, as illus-
trated in Fig. 2.2b. This highlights the relevance of a clear understanding of both
Efficiency,%
60 60 US$1.00/W
US$1.00/W
40 III 40 I II
Present limit
20 20 US$3.50/W
I US$3.50/W I
II II
0 100 200 300 400 500 200 100 0 100 200 300 400 500
Module Cost, US$/m2 System Cost, Module Cost,
US$/m2 US$/m2
(a) (b)
Figure 2.2 (a) Efficiency-cost trade-off for three generations of photovoltaic technology,
based on (I) silicon wafers, (II) ‘second-generation’ thin-films and (III) advanced, high-
efficiency, thin-film ‘third generation’ approaches; (b) Additional leverage provided by
high efficiency when additional area-related costs are included. Efficiency limits are shown
as bands to accommodate the increased efficiencies possible if sunlight is concentrated (or
the acceptance angle of the cell reduced). Source: Green and Ho (2011).
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch02 page 44
44 M. A. Green
the efficiency limits that apply to the various photovoltaic technologies and also of
the improvements required to progress towards them.
excited states
(- ) (+)
ground states
Figure 2.3 Essential requirements for photovoltaics, selectively contacted ground and
excited states. Source: Green and Hansen (2002).
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch02 page 45
carriers. What is also needed in an efficient solar cell is a way of taking up this
voltage change without affecting carrier collection.
For p-n junction devices, this mechanism is provided by the ‘built-in’ voltage
at the junction, which has a value at thermal equilibrium that depends on the product
of the doping concentration on either side of the junction. This junction potential
drop reduces under light exposure, providing the electrostatic adjustment needed
to accommodate the generation of photovoltage. If an attempt is made to exceed
the built-in voltage, the consequences depend on the conditions at the contacts. If
the doping is higher here than near the junction region, as in the mainstream silicon
wafer commercial product, additional adjustment can occur such as at the potential
step associated with the rear back surface field (BSF). The doping of a p-n junction
also provides the selective contacting requirement, with contact made selectively
to the conduction band on the n-type side of the device and to the valence band on
the p-type side of the device. The selectivity is provided by the different numbers
of conduction-band electrons and valence-band holes at the two contacts (Fig. 2.4).
For thin p-i -n structures, such as amorphous silicon devices, a similar built-
in voltage is present, corresponding to a nearly uniform field across the i -region,
with this field aiding carrier collection. This field’s reduction on illumination, as
photovoltage increases, reduces the effectiveness of carrier collection, resulting in
reduced fill factors. Doping again provides the contacting selectivity.
For dye-sensitised cells, a barrier layer at the TiO2 /front contact interface
provides a built-in potential reservoir, with an insufficient value of this potential
producing low fill factors (Kron et al., 2003; Snaith and Grätzel, 2006). Contact
to the excited states is provided by the n-type TiO2 , while contact to the ground
Transition
region
p-type n-type
Figure 2.4 Essentials of p-n junction solar cell operation. Doping provides selective contact
to the appropriate band by controlling occupancy in the contact region. Source: Green and
Hansen (2002).
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch02 page 46
46 M. A. Green
Figure 2.5 (a) Energy spectral density for the presently accepted photovoltaic reference
spectra (entropy spectral density is also shown in the lower section of the plot on an enlarged
scale as dashed lines for the direct and diluted AM0 spectrum); (b) Planckian equivalent
temperature versus wavelength for the three spectra. Source: Green and Ho (2011).
g = Aπ. In the latter case, integrating Eq. (2.1) from 0 to ∞ gives the result σ T 4
per unit area where σ , the Stefan–Boltzmann constant, equals 2π 5 k 4 /(15h 3 c3 ). A
similar expression to Eq. (2.1) applies to the photon flux Ṅ (E 1 , E 2 ) but with E 3
in the numerator replaced by E 2 .
For radiation that differs from the black-body spectrum, the same expres-
sions can still be used by defining a ‘Planckian’ temperature T (E), a function
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch02 page 48
48 M. A. Green
of energy, which gives the correct magnitude for the spectral component at each
energy. Figure 2.5b (upper curves) shows the calculated Planckian temperatures
associated with the AM0 and AM1.5 spectra. The variation in this temperature
for the AM0 spectrum is largely due to different absorption strengths in the Sun’s
photosphere for different wavelengths, resulting in the sampling of temperatures
at different depths. For the AM1.5 spectra, there is the additional impact of atmo-
spheric absorption bands, notably the very strong water vapour absorption bands
centred at about 0.95, 1.1, 1.4, 1.9 and 2.7 µm, although the latter is at too long a
wavelength to be particularly significant for solar energy conversion.
The above Planckian effective temperatures have thermodynamic significance.
If the photon occupancy number is defined as
n ph = 1/(e E/ kT − 1) (2.2)
Ts
Tsky
dA
x
y
Figure 2.6 Illustration of radiation environment seen by an element on the cell’s surface.
As well as the direct sunlight at temperature TS , the cell is also exposed to sky radiation at a
temperature Tsky assumed to equal the ambient temperature T A . Source: Green and Hansen
(2002).
50 M. A. Green
TA
.
E S .
W
.
SS
Tc
.
.
Q
. . .
EC S = Q/T A
.
S C
.S >0
G
Figure 2.7 System for calculating the Landsberg efficiency limit. Source: Green and Hansen
(2002).
where η is the energy conversion efficiency, Ẇ is the useful work (electricity) and
the Ṡ and Ė terms correspond to the associated entropy and energy fluxes. Noting
TC ≥ T A , and assuming the cell is a perfect absorber over all wavelengths emitted
by the Sun (both black bodies), the maximum value of this efficiency occurs when
ṠG = 0 and TC = T A , giving
4 TA 1 T A4
ηL = 1 − + (2.5)
3 TS 3 TS4
This limit is known as the Landsberg limit (Landsberg and Tonge, 1980).
For a 6000 K Sun and a 300 K ambient, the resulting efficiency limit is 93.3%.
While it is generally accepted that this represents an upper bound on the efficiency
of solar energy conversion, some have thought this bound cannot be reached in
principle, as previously noted, due to unavoidable entropy generation during light
absorption (making the ṠG = 0 condition fundamentally unattainable). However,
the arguments involved apply only to time-symmetric systems. It has been shown
that, in principle, the Landsberg limit can be approached arbitrarily closely by
systems that are time-asymmetric (Ries, 1983).
A subtlety is that this limit applies only to systems involving a direct exchange
with the Sun. This can be achieved by concentrating optics if sunlight is concen-
trated to the maximum extent possible (46,200 times). During such concentration,
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch02 page 51
For the present case of dilution by 46,200 times, χ has a value of 3.94, giving
a limiting Landsberg efficiency for isotropic converters of 73.7% for a 6000 K Sun
and a 300 K ambient, the value shown in Fig. 2.1.
Figure 2.8 shows a photovoltaic system capable in principle of reaching such
limits (Green, 2003). Each converter in the system consists of a stack of a large
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch02 page 52
52 M. A. Green
number of cells of different bandgap. The first converter accepts the full sunlight
intensity from the Sun via a circulator, which allows the light emitted by the cell
to be steered to the next cell. In this way, each cell can be operated infinitesimally
close to open-circuit, where it converts all wavelengths at the Carnot efficiency,
with essentially only ambient temperature black-body radiation eventually sent
back to the Sun.
Clearly, a much simplified way of implementing time-asymmetric effects will
be required if they are to be used to practical advantage. An encouraging result
is that, if only one circulator and two cell stacks are used, the maximal efficiency
increases from 86.8% to 90.7%, over half the possible benefit from time asymmetry
(Brown, 2003).
hf >> Eg hf < E g
Eg
54 M. A. Green
Hence, all orders will have roughly similar values for small z. Repeating the
previous integration for n = 3 gives
∞ 3
ε dε
ε
= εg3 Li 1 (e−εg ) + 3εg2 Li 2 (e−εg ) + 6εg Li 3 (e−εg ) + 6Li 4 (e−εg )
εg e − 1
(2.15)
Hence, the Trivich–Flinn limit expressed in polylogarithms is
εg [εg2 Li 1 (e−εg ) + 2εg Li 2 (e−εg ) + 2Li 3 (e−εg )]
ηTF = (2.16)
6Li 4 (1)
where Li4 (1) is an Euler series with value π 4 /90. The value of εg giving the max-
imum TF efficiency (εg0 ) can be found by differentiating the above numerator to
give
εg0
3
Li 0 (e−εg0 ) = εg0
2
Li 1 (e−εg0 ) + 2εg0 Li 2 (e−εg0 ) + 2Li 3 (e−εg0 ) (2.17)
This equation has the solution ε g0 = 2.1657 . . . normalised to kT S with the
corresponding maximum value of the TF efficiency of 43.88%, regardless of TS
(the assumed black-body temperature of the Sun). In terms of εg0 , it can be seen
by comparing Eqs (2.16) and (2.17) that this maximum efficiency is given by
15 εg0
4
ηTF
m
= (2.18)
π 4 eεg0 − 1
For a 6000 K black-body source, the optimum bandgap would be 1.120 eV,
about the bandgap of silicon.
that a strongly absorbing cell in thermal equilibrium with its ambient at temper-
ature T A would be emitting black-body radiation. For photon energies above the
cell bandgap, the physical source of the emitted photons would be band-to-band
radiative recombination, the inverse of the absorption process of Fig. 2.9.
In a non-equilibrium situation with voltage across the device, SQ were aware
that, in a high-quality p-n junction device, the electron–hole product, and hence
these radiative recombination rates, would be exponentially enhanced throughout
the cell volume by the cell voltage, V , normalised to the ‘thermal voltage’, kTA /q.
Hence, a good-quality cell, for photon energies above its bandgap, would emit
exponentially enhanced black-body radiation. Assuming 100% quantum efficiency,
one electron would flow into the cell terminals for each emitted photon.
Using this insight, but a more general chemical potential model of photon
emission (Würfel, 1982), the electrical power density output from a photovoltaic
cell in the SQ limit can be expressed as (Green, 2003)
PSQ = V J
2πq V ∞ E 2d E ∞ E 2d E
= 3 2 ξ −
h c Eg exp(E/kTS ) − 1 Eg exp[(E − q V )/kTA ] − 1
(2.19)
56 M. A. Green
where γ = e(q V −E g )/ kTA and t = TA /TS . Comparison with Eq. (2.16) shows the
TF formulation is a limiting case of the SQ formulation for TA = 0. This close
relationship will be exploited in the following.
Unlike Eq. (2.16), Eq. (2.21) is a function of not one but three variables εg ,
V and t. Since these variables can be expressed in terms of standard functions
supported by common mathematical packages such as Mathematica (2008), even
in this form Eq. (2.21) could be simply solved, for example, for its global maximum
value at fixed t (built-in Mathematica programs such as FindMaximum require
only the entry of the equation, plus essentially a single line of code, to find the
solution).
The technique generally used to find the peak SQ efficiency is to fix E g , find
the optimum V for this E g , and hence steadily map out efficiency as a function of
E g (Shockley and Queisser, 1961). However, partial differentiation of Eq. (2.19)
or (2.21) with respect to E g locates a stationary point when
For the case of undiluted sunlight (ξ = 1), this simplifies to the Carnot rela-
tionship noted elsewhere (Landsberg, 2000)
ξ =1 15
ηSQLB = εg (1 − t)2 [εg2 Li 1 (e−εg )
π4
+ 2(1 + t)εg Li 2 (e−εg ) + 2(1 + t + t 2 )Li 3 (e−εg )] (2.24)
εgm
3
Li 0 (e−εgm ) = (1 − 2t)εgm
2
Li (e−εgm ) + 2(1 + t − t 2 )εgm Li 2 (e−εgm )
where since εgm is in normalised units, the actual bandgap E gm = εgm kTS .
Noting the TF formulation is a limiting case of the SQ formulation for TA = 0
and that small TA /TS ratios (∼0.05) prevail for solar conversion, the SQ results
can be expressed linearly in terms of TF results. The exact value of the linear
term in a power series expansion can be found by solving Eq. (2.21) by a single
Newton–Raphson iteration, using εg0 as the trial solution, giving
ξ =1
εgm (SQ) = εg0 − t[3Li 1 (e−εg0 ) − εg0 Li 0 (e−εg0 )]/
[4Li 0 (e−εg0 ) − εg0 Li −1 (e−εg0 )]
= 2.1657 − 0.422 TA /TS (2.26)
Both expressions are extremely accurate, giving the correct values of bandgap
and efficiency for t = 0.05 to within one digit in the fifth and fourth significant
figure, respectively.
For the case of diluted sunlight, of considerable practical importance since
it applies to the general photovoltaic case of non-maximal sunlight concentration
and also, as discussed below, to cases where non-radiative recombination occurs
in parallel with the modelled radiative recombination, Eq. (2.19) again prevails at
the optimum V . At this optimum, γ depends only on εg and ξ with a value γ ∗
equal to or slightly higher than ξ e−εg as defined below
ξ
γ∗ = (2.28)
eε g −1+ξ
Proceeding as before, this gives the lower bound on SQ efficiency as
15
ηSQLB = [εg + t ln(γ ∗ )]{εg2 [Li 1 (e−εg ) − t Li 1 (γ ∗ )/ξ ]
π4
+ 2εg [Li 2 (e−εg ) − t 2 Li 2 (γ ∗ )/ξ ]
+ 2[Li 3 (e−εg ) − t 3 Li 3 (γ ∗ )/ξ ]} (2.29)
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch02 page 58
58 M. A. Green
Differentiating Eq. (2.29) again provides the relationship for the optimum
bandgap, εgm (SQ) normalised to kT S . Maximum efficiency is obtained when the
bandgap satisfies the following relationship
Figure 2.10 Isotropic performance limits for cells under the AM0, AM1.5G and AM1.5D
spectral as well as for the AM1.5D spectra, under various concentration levels. The Trivich–
Flinn limit for the AM1.5D spectrum is shown for comparison. Source: Green and Ho (2011).
approach. The uppermost curve is the Trivich–Flinn limit for the AM1.5D spec-
trum, which is approached for small-bandgap cells under high concentration levels
(or, more generally, severe angular restriction of the converter response). Note that
the limiting efficiency under concentration (or angular restriction) increases more
rapidly for small-bandgap cells than for large-bandgap cells (simply because the
approximately 60 mV increase in open-circuit voltage for ten times increase in
intensity, or decrease in angular acceptance, is proportionately larger for small-
bandgap cells).
For the AM0 spectrum for a cell with isotropic response, there is a single peak
(30.4% efficiency for E g = 1.245 eV). This is similar to the value of 31.0% at a
bandgap of 1.31 eV calculated for a 6000 K black-body spectrum at a cell temper-
ature of 300 K. For the terrestrial AM1.5 spectra, there are multiple peaks due to
the strong water vapour absorption bands apparent in Fig. 2.5. The most impor-
tant are the absorption bands in the 920–970 nm (1.28–1.35 eV), 1100–1160 nm
(1.07–1.13 eV) and 1300–1500 nm (0.83–0.95 eV) wavelength (energy) ranges.
Loss of radiation in the terrestrial spectra due to the first of these bands elimi-
nates the ‘natural’ peak apparent in the AM0 spectrum, creating a local maximum
on the high-energy side of the absorption band and a local minimum on the low-
energy side.
Note that from Fig. 2.10 the AM0 and AM1.5G spectra give relatively higher
efficiency for larger bandgaps than does the AM1.5D spectrum. This is due to blue
photons being selectively scattered out of the AM1.5D spectrum during passage
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch02 page 60
60 M. A. Green
Figure 2.11 Isotropic efficiency limits under the AM1.5G spectrum, and also for devices
with 1% and 0.01% radiative efficiency. Also shown are best-certified cell limits efficiencies
for various cell technologies. Source: Green and Ho (2011).
through the atmosphere. Combined with the intensity dependence already noted,
this shifts the global maximum between three different peaks, each associated with
the high-energy side of the three absorption bands already mentioned.
The solid line of Fig. 2.11 shows the limiting efficiency for a conventional cell
with isotropic response under the AM1.5G reference spectrum. Compared with an
earlier spectrum in use until 2008 (Green et al., 2009), the stronger spectral content
over the 650–900 nm range shifts the peak efficiency to higher energy, so that the
two peaks around 1.15 eV and 1.35 eV are nearly equal, with peak efficiencies of
33.6% and 33.8% respectively. For the previous reference spectrum (ASTM E892-
87, IEC60904-3:1989), the order of the peaks was reversed and both were about
1% lower in efficiency (Tiedje et al., 1984), consistent with expectations discussed
elsewhere (Green et al., 2009).
Also shown in Fig. 2.11 as dashed lines are the limiting efficiencies for 0.01%
and 1% radiative efficiencies of the cell, with radiative efficiency defined as the
fraction of nett recombination in the device that is radiative. Radiative efficiency
is a measure of the state of development of the cell material technology. Limiting
efficiencies calculated with reduced radiative efficiency are identical to those cal-
culated under a reduced intensity (further dilution) of sunlight. Hence the previous
curves under concentration (Fig. 2.10) can also be used to calculate the effects
of non-ideal radiative efficiency in this case as well (for example, the curve for
100 kW m−2 incident intensity represents the limiting performance for a cell under
1000 kW m−2 intensity, if the radiative efficiency is 10% rather than 100%).
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch02 page 61
Given this correspondence, it is not surprising that the efficiency for low-
bandgap cells is most sensitive to low values of the radiative efficiency, since a
60 mV/decade voltage reduction with decreasing radiative efficiency has a propor-
tionately more devastating impact. Peak efficiency is pushed to higher-bandgap
cells as the radiative efficiency decreases.
Also shown in Fig. 2.11 are best-certified experimental values for a range
of different solar cell technologies (Green et al., 2011). ‘Bandgap’ in this case
corresponds to the energy on the low-energy side of the cell’s spectral response
curve, where external quantum efficiency (EQE) reaches 50% of its peak value.
Silicon (c-Si) and c-GaAs lie above the 0.01% radiative efficiency line. Actual
radiative efficiencies are close to or well above 1% for both technologies. The
difference arises since experimental devices have additional losses other than the
radiative loss assumed including, for example, reflection and resistance losses, but
also including additional radiative loss due to non-abrupt absorption thresholds
(Kirchartz et al., 2009).
62 M. A. Green
do not emit radiation above the cell bandgap exactly as assumed in the Shockley–
Queisser analysis (exponentially enhanced by the cell voltage with the spectral
distribution of room-temperature black-body radiation). Rather, they emit radia-
tion with this distribution multiplied at each wavelength by the external quantum
efficiency (EQE) of the device, as measured for the device operating as a solar cell.
This remarkable relationship has been confirmed for a number of experimental
devices (Kirchartz and Rau, 2007; Kirchartz et al., 2008a; Kirchartz et al., 2009;
Vandewal et al., 2009), as well as being consistent with modelling prior to its
recognition (Green et al., 2001).
The cell EQE must be measured for a calibrated measurement of cell effi-
ciency, to adjust for spectral mismatch between the spectrum used for illuminating
the cell during measurement and the tabulated reference spectrum (International
Standard, 2008). This means that each calibrated cell measurement yields sufficient
information to allow calculation of ERE.
On cell open circuit, the photogeneration of carriers within a cell is bal-
anced by recombination within the device, globally if not at each point within
the device. The open-circuit voltage measured at the cell terminals is the voltage
that increases recombination within the device-active region to the level required
to balance the photocurrent able to be collected by the junction. Assuming reason-
able device quality, so parasitic resistances are not excessive, and also assuming
that the collection probability of carriers is not strongly voltage-dependent, the
photocurrent density collected by the junction on open circuit will equal the short-
circuit current density. The light emitted by the junction will be the EQE-weighted,
exponentially-enhanced black-body radiation described above, allowing the ERE
to be calculated as
2πq q Voc E Q E abs E 2 dE q Voc
h 3 c2
exp kT exp(E/ kT )−1 exp kT E Q E rel NBB (E)d E
ERE = =
Isc /A E Q E rel N AM1.5 (E)d E
(2.33)
EQE is the value for the solar cell for near perpendicularly incident light in
either absolute or relative terms, with EQE the Lambertian weighted value over all
angles of incident light, of absolute value EQEabs and relative value EQErel . For
a high-quality cell, this will not differ greatly from the near perpendicular value.
EQE is typically measured either in absolute or relative terms with a midrange
accuracy of about 3% relative. The major contributions to the integration on the
numerator come from the long-wavelength region of the spectral response, from
the region where this response begins to fall rapidly to the low value near the
absorption threshold, while the main contribution to the denominator comes from
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch02 page 63
the region near the peak photon density in the AM1.5 spectrum, which is in the
600–700 nm range.
Some of the sources of inaccuracies in the evaluation of ERE have already been
suggested. An isotropic response needs to be assumed for evaluation from standard
near-perpendicular EQE data. Collection probabilities may be voltage-dependent
so that the short-circuit condition, where EQE is normally evaluated, may not
represent conditions at open circuit. The reciprocal relation itself is also only strictly
valid under conditions where the quasi-Fermi level separation at the junction is
constant (Rau, 2007), which would not be the case for resistive devices even at
open circuit, due to circulating currents. However, these effects are expected to be
minor for cells of respectable performance, particularly compared with the large
differences in ERE noted between different devices and different technologies.
ERE was calculated for a range of representative state-of-the-art cells (Green
et al., 2010; Green et al., 2011). The absolute EQEs for these cells are shown
in Fig. 2.12a, deduced from the relative EQE by normalising to the experimen-
tal short-circuit current density measured for the devices. The calculated spectral
luminescence from these devices on open-circuit is shown in Fig. 2.12b. Units in
this case are A/m2 /nm, representing the current density required on open circuit to
support the different spectral emission components shown. Due to large differences
between technologies, different scaling factors are applied to the different results,
although the same scaling factor is applied to cells of the same type, except for the
GaAs cells.
Until recently (Green et al., 2011) the most efficient non-concentrating, single-
junction cell was a 26.4% efficient GaAs cell fabricated by the Fraunhofer Institute
for Solar Energy (GaAs ISE), with parameters shown in Table 2.1 Integrating
the calculated spectral luminescence gives an ERE of 1.26%. Given that this device
was fabricated on a GaAs substrate, IRE would have been appreciably higher since
n 2 more radiation (where n is the substrate refractive index) would have been
emitted into the substrate (with most absorbed there) than into air.
In late 2010, this record was surpassed with 27.6% efficiency measured for a
thin-film GaAs device fabricated by Alta Solar (GaAs Alta). As shown in Fig. 2.12a,
this has an almost identical spectral response to cell GaAs ISE but, as in Fig. 2.12b,
the luminescent intensity on open circuit is almost 20 times higher, corresponding
to a greatly improved ERE of 22.5%. This high value suggests photons emitted
during radiative recombination events in a direction towards the rear of the device
are not wasted as in the GaAs ISE device, but that a reasonable number are reflected
back into active device regions.
With radiative recombination forming such a large fraction of total recombina-
tion in this device, further improvements in ERE will result in immediate benefits.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch02 page 64
64 M. A. Green
Figure 2.12 (a) EQE for the different cells described in the text; (b) calculated spectral
luminescence from the different cells on open circuit, with widely varying multiplication
factors. Source: Green and Ho (2011).
the present approach. Since the Sunpower device (Si SPWR) with less than 200 µm
thickness is appreciably thinner than the University of New South Wales device
(Si UNSW) with over 400 µm thickness and has possibly worse rear reflection,
the EQE is notably poorer at long wavelengths, with i sc also appreciably lower.
This is compensated by a higher Voc , due to the smaller volume of the bulk region,
probably combined with better surface passivation. From Fig. 2.12b and Table 2.1,
the two devices have almost identical open-circuit voltage luminescence.
Comparison of two recent record CIGS (copper indium gallium selenide)
devices also produces interesting results. One is a small area 20.3%-efficient device
fabricated by Zentrum für Sonnenenergie-und-Wasserstoff-Forschung (CIGS
ZSW), and the second is for a larger (1 cm2 ) device of 19.6% efficiency fabri-
cated by the US National Renewable Energy Laboratory (CIGS NREL). From
Fig. 2.12a, the two devices can be seen to have similar spectral responses, although
the CIGS ZSW device has a much stronger emission on open circuit, with ERE
of 0.19% compared to 0.06% for the NREL device. This indicates fundamen-
tally better quality material in the ZSW device, although this may be partly due
to its smaller size, given the strong dependence of CIGS cell performance on
cell area.
The next three devices discussed include recent record CdTe, a-Si and dye-
sensitised devices. The CdTe device is a small 12.5%-efficient submodule with
ERE of 1.0 × 10−4 %. Smaller cells could be expected to have higher ERE while
commercial CdTe modules, averaging about 11.4% aperture area efficiency in
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch02 page 66
66 M. A. Green
2011, would be expected to have lower ERE. The best a-Si cell to date has quite a
low ERE of 5.3 × 10−6 %. However, from Fig. 2.11, its bandgap can be seen to be
in an appropriate region for the highest possible conversion efficiency with such
low ERE.
The dye-sensitised device is again a small 9.9% efficient submodule, although
with energy conversion performance close to the best-performing small-area cell
(11.2%). Although the operational principles of a dye-sensitised cell are vastly dif-
ferent from the previous p-n junction devices, the reciprocity relationships giving
rise to Rau’s relationship are expected still to apply (Trupke et al., 1999). ERE is
7.2 × 10−6 %, similar to, but slightly higher than, the a-Si device.
The final two devices are two of the first organic photovoltaic devices to exceed
8% energy conversion efficiency, with very different energy absorption thresholds,
as apparent from Fig. 2.12a. Both devices, however, display very similar ERE in
the 3−4 × 10−7 % range. A detailed discussion of the electroluminescence from a
range of organic cells is presented elsewhere (Vandewal et al., 2009).
Even though organic light-emitting diodes (OLEDs) of quantum efficiency
>20% have been reported, the different requirements for photovoltaics (Kirchartz
et al., 2008b) greatly reduce the radiative efficiency. In particular, the blending
required to produce bulk heterojunction devices with high carrier collection greatly
reduces the radiative efficiency.
The above results show that the external radiative efficiency (ERE) can be
unambiguously deduced from standard solar cell efficiency measurements and is
a useful parameter in comparing the performance of cells of both the same and
completely different technologies. As each technology matures, ERE will evolve
towards the 100% value required for limiting performance.
N n+1 n n-1 1
Figure 2.13 Optimal configuration for stacked tandem cells. After Marti and Araujo (1996).
2πqξ Vn
Pn = {ε2 [Li 1 (e−εgn ) − t Li 1 (γn )] + 2εgn [Li 2 (e−εgn ) − t 2 Li 2 (γn )]
h 3 c2 (kTs )3 gn
+
+ 2[Li 3 (e−εgn ) − t 3 Li 3 (γn )] − εgn
+2
[Li 1 (e−εgn ) − t Li 1 (γn+ )]
+ +
+
− 2εgn [Li 2 (e−εgn ) − t 2 Li 2 (γn+ )] − 2[Li 3 (e−εgn ) − t 3 Li 3 (γn+ )]} (2.35)
+
where γn = e(q Vn −E gn )/ kTA and γn+ = e(q Vn −E gn )/ kTA .
With the equations formulated in this way in terms of standard functions,
global maxima can be simply found with standard mathematic packages just by
entering these equations constrained by current matching if this feature is desired.
Figure 2.14 was generated for the unconstrained 3-cell case with a simple contour
plot command using Mathematica, involving essentially two steps of code.
68 M. A. Green
Figure 2.14 Contour plots for 47%, 48% and 49% efficiency versus normalised bandgap
εg1 , εg2 and εg3 , for a 3-cell tandem stack (unconstrained) with t = 0.05 and ξ = 1/46,200.
Peak efficiency is 49.3% for εg1 = 1.58, εg2 = 2.77 and εg3 = 4.36 (multiply by 0.517 eV
to convert to actual bandgap values for TA = 300 K, TS = 6000 K). Source: Green and Ho
(2011).
Partial differentiation with respect to V gives the following two equations for
the optimal value Vm
q Vm
Li 0 (e(q Vm −E g )/ kTA ) + Li −1 (e(q Vm −E g )/ kTA )
kTA
= ξ Li 0 (e−E g / kTs ) + (1 − ξ )Li 0 (e−E g / kTA )
= Li 0 (eq(Voc −E g )/ kTA ) (2.38)
q Voc = E g − kTA ln{1 + 1/[ξ Li 0 (e−E g / kTs ) + (1 − ξ )Li 0 (e−E g / kTA )]}
(2.39)
and
kTA q Voc
Vm∗ = Voc − ln /(1 − γoc) + 1 (2.43)
q kTA
Since the term in the square bracket is already an overestimate, deleting the
(1 − γm ) and the (+1) terms will generally improve accuracy, provided Voc is
constrained to values greater than kT A . Using this modified version of Eq. (2.43)
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch02 page 70
70 M. A. Green
in combination with the asymptotic expansion of Eq. (2.41) gives an explicit rela-
tionship for Vm∗ which remains useful for E g ≥ kTA [1 − ln(ξ )]/(1 − TA /Ts ). For
smaller E g , only a small power contribution would be generated, so setting Vm to
zero for such E g will not cause any significant error in total output. This approxi-
mation was found to result in efficiencies which matched exact values to close to
three significant digits over the whole range of likely interest.
The resulting expression may also be integrated analytically. Inserting into
Eq. (2.37) and expanding to first order in TA , then integrating, gives the result
where
(a) (b)
Figure 2.15 Normally photogenerated carriers thermalise with the lattice as in (a). In hot-
carrier cells, excess energy is stored in a hot-carrier distribution, as shown in (b). Source:
Green and Hansen (2002).
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch02 page 72
72 M. A. Green
slow the decay of optical into acoustic phonons, and hence the energy relaxation,
of carriers by reducing the rate of energy exchange with optical phonons.
Calculations involving the relevant experimental parameters of quantum wells
and bulk materials show that hot-carrier effects should be apparent in suitably
designed devices operating under highly concentrated sunlight (Guillemoles et al.,
2005; Aliberti et al., 2010). Development of new materials will be required to
extend this attribute to devices operating under global sunlight.
IQE
3.0
2.0
1.0
0 λ
λG
(a) (b)
Figure 2.16 (a) Schematic of the impact ionisation process whereby one energetic photon
creates multiple electron–hole pairs; (b) energetically feasible internal quantum efficiency,
where λg is the wavelength corresponding to the threshold energy for electron–hole pair
creation. Source: Green and Hansen (2002).
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch02 page 73
86.8% for a 6000 K black-body source). However, MEG device are less suited to
global sunlight, with the limiting efficiency dropping considerably, to just above
that of a conventional 2-cell tandem stack (43.6% for a 6000 K black-body source,
compared with 42.9% for a 2-cell stack and 68.2% for the time-symmetric limit).
This is due to the larger bandgap required for efficient operation at low intensities,
reducing the gain from multiple carrier generation (Green, 2003). However, this
approach may well provide a way of improving cell performance with specific cell
technologies.
conduction
band
impurity
3 subgap level
photons
1
valence
band
Figure 2.17 Impurity photovoltaic effect where electron-hole pairs are generated by sub-
bandgap photons. After Keevers and Green (1994).
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch02 page 74
74 M. A. Green
C2
C1
C3
Figure 2.18 Equivalent circuit of the intermediate-band solar cell. The circuit elements
shown represent ideal Shockley–Queisser solar cells. Source: Green and Ho (2011).
The multiple quantum well cell (Barnham and Duggan, 1990; also Chapter
10 of this volume) and quantum-dot structures (Luque and Marti, 2010) have
become the preferred devices for investigating the implementation of this approach.
Although devices with creditable performance have been demonstrated, these have
not yet exceeded the performance of controls without the intermediate levels.
Incident light
C3
C1 C2
Solar cell
Insulator
C4
Reflector
(a) (b)
Figure 2.19 (a) Solar cell and an electronically isolated up-converter. Sub-band-gap light
transmitted by the solar cell is up-converted into high-energy photons, which are subse-
quently absorbed in the solar cell; (b) equivalent circuit of the up-conversion system. The
solar cell in front of the up-converter is denoted C1. The up-converter is represented by the
three cells C2, C3 and C4. The series connected cells C3 and C4 represent the two inter-
mediate transitions while C2 represents the band-to-band transitions. After Trupke et al.
(2002a).
Incident light
Incident C3
C2 C1
Converter
Insulator C4
Solar cell
Reflector
(a) (b)
Figure 2.20 (a) Schematic diagram of the down-conversion system with the luminescence
converter located on the front surface of a solar cell. High-energy photons with h̄ν > 2E g
are absorbed by the converter and down-converted into two lower energy photons with
h̄ν > E g , which can both be absorbed by the solar cell; (b) equivalent circuit of the system.
The luminescence converter is represented by three solar cells C2, C3, and C4 representing
the band-to-band transitions and the two types of intermediate transitions, respectively.
High-energy photons are absorbed by C2, which then biases the cells C3 and C4 into the
forward direction and causes them to emit electroluminescence absorbed by the main cell
C1. After Trupke et al. (2002b).
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch02 page 76
76 M. A. Green
W
TR
Sun
QT TA
R
Carnot
converter
QT
A
re-emitted
Figure 2.21 Solar thermal system based on the conversion of heat collected by an absorber
to electricity at the Carnot efficiency. Source: Green and Hansen (2002).
can be calculated from the schematic of Fig. 2.21. A black absorber receives energy
from the Sun, heating it to a high temperature, TR . Surfaces of the absorber other
than those receiving energy from the Sun have low emissivity so that only the
receiving surface loses energy back to the Sun. Heat from the absorber drives a
Carnot converter.
The limiting efficiency can be written down almost by inspection for a black-
body Sun
TR4 TA
ηT = 1 − 1− (2.46)
ξ TS4 TR
Figure 2.22 Limiting efficiency for the scheme of Fig. 2.21. Source: Green and Ho (2011).
As seen by the upper line in Fig. 2.22, use of such a selective absorber
significantly improves the energy conversion efficiency at low incident sunlight
intensities.
Similar energy conversion efficiency limits to those shown in Fig. 2.22 apply to
photovoltaic devices that follow a thermal route to conversion. In considering these
results, it should be borne in mind that these thermal limits are much more difficult
to attain in practice than the corresponding limits for conventional photovoltaic
devices, since the latter already take into account some of the practical difficulties
in obtaining Carnot-like efficiencies.
A solar thermophotovoltaic system uses the radiant energy from the receiver
rather than its heat energy. Figure 2.23 shows a schematic illustrating the basic
concept. An ideal solar cell with an ideal monochromatic light filter acts as an
ideal Carnot converter of radiation emitted by the receiver (Green, 2003). Hence
the scheme of Fig. 2.23 fulfils the ideal and is bounded by the same limiting
efficiencies as for Fig. 2.22b.
High expectations were held for this scheme in the late 1970s (Swanson, 1979),
but results to date have been quite modest. A good explanation of the reasons why
one might expect lower than ideal performance is given by Harder and Würfel
(2003).
A development of the idea is based on the realisation that a light-emitting diode
can convert electricity to light with greater than 100% power conversion efficiency,
if heat input to the diode is ignored. This gives rise to the ‘thermophotonic’ converter
of Fig. 2.24, which has advantages in terms of the narrow range of emission energies
of the heated device and much stronger emission at low receiver temperatures.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch02 page 78
78 M. A. Green
sunlight
NP
Figure 2.23 Thermophotovoltaic conversion with narrow passband filter. Source: Green
and Hansen (2002).
IH IC
filter
+ +
VH VC
NP PN
TH TC
Figure 2.24 Thermophotonic conversion, with narrow-pass filter and independent power
supplies. V and I are the voltages and currents at the terminals of the p-n junction devices,
while T is the device temperature and the subscripts H and C refer to the hot device heated
by sunlight and the cold receiver, respectively. Source: Green and Hansen (2002).
electrons
anode
cathode
sunlight
electricity
Figure 2.25 Thermionic conversion. Sunlight heats a cathode that emits electrons which
are collected by a lower work function anode.
been very modest (well below 1% and even 0.1%). Radiative losses to the cathode
(despite attempts to reduce the emissivity of the facing surfaces of the anode and
cathode), conduction losses through wiring to the electrodes, resistive losses in the
wires and space-charge effects are key contributors to the difference (Hatsopoulos
and Gyftopoulos, 1979; Habedank, 1993).
It has been suggested that efficiency can be improved by combining both
heat and light excitation of the cathode. In one case, combined photoemission and
thermionic emission has been suggested, with the latter possibly augmented by
hot-carrier effects (Smestad, 2004). A more recent suggestion has been to use a
p-type semiconductor as the heated cathode, with above-bandgap light used to
photoexcite the cathode and sub-bandgap light and carrier relaxation losses also
used to heat the cathode (Schwede et al, 2010).
The advantage of the latter configuration is that the flatband condition is still
determined by the relatively high work function of the p-type semiconductor, estab-
lished by the majority-carrier Fermi level. However, photoexcitation moves the
minority-carrier electron quasi-Fermi level to closer to the vacuum level, enhancing
photoemission (and also reducing the minimum value of the anode work function).
The disadvantage is that to absorb above-bandgap light, the anode must also emit
it back to the Sun. For 100% radiative efficiency, this means each absorbed pho-
ton will result in a photon of energy just above the bandgap being emitted back
to the Sun, a Trivich–Flinn inefficiency. For some bandgaps, this will represent
40% of the incident energy, a huge loss before any serious attempt at conversion
begins.
Since the Carnot conversion potential of standard thermionic conversion obvi-
ously cannot be exceeded in this way, the advantage lies in offering lower possible
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch02 page 80
80 M. A. Green
operating temperatures for the converter, which may allow some of the previous
losses mentioned in traditional converters to be reduced.
The other traditional converter of heat to electricity is a thermoelectric con-
verter. The close relationship to photovoltaics is highlighted by a casual remark by
Würfel (2000) that one way of reducing heat loss to the lattice in a hot-carrier cell
is to let the lattice heat up to the hot-carrier temperature. A heated absorber with
selective energy contacts (Section 2.7.1) acts as a hot-carrier cell in the Würfel
limit (Würfel, 1995; Green, 2003). The conversion efficiency limits upon such a
cell are identical to the limits plotted in Fig. 2.22. This type of device has been
independently suggested as an improved thermoelectric converter (O’Dwyer et al.,
2005), demonstrating the close links of thermoelectrics to advanced photovoltaic
concepts.
2.8 Summary
Although most commercial single-junction solar cells operate at more than 50%
of their theoretical potential, there is scope for improving performance by edging
closer to this limit. Even more scope for performance gains arises through advanced
‘third generation’ options that overcome the basic trade-off in a standard cell: too
high a bandgap results in too little current; too low a bandgap gives too little
voltage.
Extension of the Shockley–Queisser approach offers a ‘bottom-up’ method for
analysing the ultimate potential of any new scheme, such as the recently suggested
photo-enhanced, thermionic conversion (Schwede et al., 2010). Formulation in
terms of polylogarithms, as in the present chapter, simplifies the programming
of the appropriate analyses. ‘Top-down’ thermodynamic analyses have already
proved useful in identifying the features required for the ultimate photovoltaic
converter. Calculations in terms of a black-body Sun have the virtue of timelessness
and also of evaluation of results analytically. However, results for the tabulated
reference spectra used for experimental cell measurements are also of considerable
interest, although these reference spectra are more ephemeral. For example, a
recent downscaling of the likely value of the solar constant from 1366.1 W m−2 to
1360.8 W m−2 (Kopp and Lean, 2011) is likely to have a flow-on effect to first the
standard AM0 spectra, then possibly the terrestrial spectra. Some key results for
present spectra of interest are compared in Table 2.2.
Apparent from Table 2.2 is the large difference between the limiting perfor-
mance of standard single-junction cells and the ultimate limits on solar energy
conversion. Capitalising on the potential factor of two performance gains would
see thin-film module technologies go beyond 25% energy conversion efficiency,
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch02 page 81
Table 2.2 Summary of photovoltaic efficiency limits under black-body radiation at 300 K
and under the three ASTM G173-03 spectra at 25 ◦ C.
Blackbody
(T S = 6000 K,
T A = 300 K) AM0 AM1.5G AM1.5D
and possibly beyond 30% in the longer term. For wafer-based technology, the
figures would be correspondingly higher.
Several options for obtaining such efficiency improvements have been
reviewed. Tandem cell stacks such as those discussed in Chapter 7 of the current vol-
ume are clearly the most promising route based on experience to date. Hot-carrier
cells have the attraction of offering higher efficiency than a 5-cell tandem stack in a
simple 2-terminal configuration, although considerable theoretical and experimen-
tal progress will be required before any performance improvement is likely to be
demonstrated with this approach. Up- and down-conversion offer the prospects for
‘supercharging’ the performance of existing cell technologies, although improve-
ments in the performance for both types of converters are required also before
any performance gain would be expected. Intermediate-band and multiple exciton
generation devices offer improvements for specific material systems.
The ‘thermal approaches’ — thermophotovoltaics, thermionics, thermoelec-
tric or hot lattice ‘hot-carrier’cells — offer tantalisingly high limiting efficiencies.
However, the gap between what is likely to be obtained in practice and the theoret-
ical potential is expected to be much larger with such approaches, because of the
much higher energy losses that are likely to occur in practice. Such approaches may
have potential in concentrating systems, but this is where tandem cell stacks have
had the most impact terrestrially and are showing unabated potential for further
development.
As is already becoming apparent as the photovoltaic industry matures, effi-
ciency will be a key driver to future cost reduction. Evidence for this is the emphasis
given by thin-film companies such as First Solar to quarterly improvements in mod-
ule efficiency (from under 8% in 2003 to over 13% in 2013) and the increasing
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch02 page 82
82 M. A. Green
Acknowledgement
This work has been supported by the Australian Government through the Australian
Renewable Energy Agency (ARENA). Responsibility for the views, information
or advice expressed herein is not accepted by the Australian Government.
References
Abengoa (2011), www.abengoasolar.com.
Aliberti P. Feng Y., Takeda Y., Shrestha S. K., Green M. A. and Conibeer G. (2010), ‘Inves-
tigation of theoretical efficiency limit of hot carriers solar cells with a bulk indium
nitride absorber’, J. Appl. Phys. 108, 94507–94517.
ASTM Standard Extraterrestrial Spectrum Reference E-490-00 (tabulated at
http://rredc.nrel.gov/solar/spectra/am0/).
Barnham K. and Duggan G. (1990), ‘A new approach to high-efficiency multi-band-gap
solar cells’, J. Appl. Phys. 67, 3490–3493.
Blom W. M., Mihailetchi V. D., Koster L. J. A. and Markov D. E. (2007), ‘Device physics
of polymer:fullerene bulk heterojunction solar cells’, Adv. Mater. 19, 1551–1566.
Brown A. S. (2003), ‘Ultimate efficiency limits of multiple energy threshold photovoltaic
devices’, Thesis for the Degree of Doctor of Philosophy, Centre for Third Generation
Photovoltaics and the School of Electrical Engineering, University of New South Wales.
Brown A. S. and Green M. A. (2002), ‘Impurity photovoltaic effect: fundamental energy
conversion efficiency limit’, J. Appl. Phys. 92, 1329–1336.
Brown A. S., Green M. A. and Corkish R. (2002), ‘Limiting efficiency for multi-band solar
cells containing three and four bands’, Physica E 14, 121–125.
Chapin D. M., Fuller, C. S. and Pearson G. L. (1954), ‘A new silicon p-n junction photocell
for converting solar radiation into electrical power’, J. Appl. Phys. 25, 676–677.
Cooke D., Gleckman P., Krebs H., O’Gallagher J., Sage D. And Winston R. (1990), ‘Sunlight
brighter than the Sun’, Nature 346, 802.
Cummerow R. L. (1954a), ‘Photovoltaic effect in p-n junctions’, Phys. Rev. 95, 16–21.
Cummerow R. L. (1954b), ‘Use of silicon p-n junctions for converting solar energy to
electrical energy’, Phys. Rev. 95, 561–562.
Deb S. and Saha H. (1972), ‘Secondary ionisation and its possible bearing on the perfor-
mance of a solar cell’, Solid State Electronics 15, 89–1391.
Green M. A. (1982), ‘Accuracy of analytical expressions for solar cell fill factors’, Solar
Cells 7, 337–340.
Green M. A. (2011), ‘Limiting photovoltaic efficiency under new ASTM G173-based
reference spectra’, Progr. Photovoltaics 20, 954–959.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch02 page 83
84 M. A. Green
Kirchartz T., Helbig A., Reetz W., Reuter M., Werner J. H. and Rau U. (2009), ‘Reciprocity
between electroluminescence and quantum efficiency used for the characterization of
silicon solar cells’, Progr. Photovoltaics 17, 394–402.
Kirchartz T., Mattheis J. and Rau U. (2008b), ‘Detailed balance theory of excitonic and bulk
heterojunction solar cells’, Phys. Rev. B 78, 235320.
Kirchartz T. and Rau U. (2007), ‘Electroluminescence analysis of high efficiency Cu(In,
Ga)Se2 solar cells’, J. Appl. Phys. 102, 104510.
Kolodinski S., Werner J. H., Wittfchen J. H. and Queisser H. J. (1994), ‘Quantum efficiencies
exceeding unity in silicon leading to novel selection principles for solar cell materials’,
Solar Energy Mater. Solar Cells 33, 275–286.
Kopp G. and Lean J. L. (2011), ‘A new, lower value of total solar irradiance: Evidence and
climate significance’, Geophys. Research Lett. 38, L01706.
Kron G., Rau U. and Werner J. H. (2003), ‘Influence of built-in voltage on the fill factor of
dye-sensitised solar cells’, J. Phys. Chem. B 107, 13258–13261.
Landsberg P. (2000), ‘Efficiencies of solar cells: where is Carnot hiding?’, Conf. Record,
16th European Photovoltaic Solar Energy Conf., Glasgow.
Landsberg P. T. and Tonge G. (1979), ‘Thermodynamics of the conversion of diluted radi-
ation’, J. Phys. A 12, 551–562.
Landsberg P. T. and Tong G. (1980), ‘Thermodynamic energy conversion efficiencies’, J.
Appl. Phys. 51, R1–R20.
Lee K., Kim J. Y. and Heeger A. J. (2008), ‘Titanium oxide films as multifunctional com-
ponents in bulk heterojunction ‘plastic’ solar cells’, in Brabecc, Dyakonovv, Scherfu
(eds), Organic Photovoltaics, Wiley, Weinheim.
Lewin L. (1981), ‘Polylogarithms and associated functions’, New York: North-Holland (also
see http://en.wikipedia.org/wiki/Polylogarithm).
Loferski J. J. (1956), ‘Theoretical considerations governing the choice of the optimum
semiconductor for photovoltaic energy conversion’, J. Appl. Phys. 27, 777–784.
Luque A. and Marti A. (1997), ‘Increasing the efficiency of ideal solar cells by photon
induced transitions at intermediate levels’, Phys. Rev. Lett. 78, 369.
Luque A. and Marti A. (2010), ‘The intermediate band solar cell: progress toward the
realization of an attractive concept’, Adv. Mater. 22, 160–174.
Luque A., Ruiz J. M., Cuevas A., Eguren J. and Agost J. G. (1978), Proc. 1st Commis-
sion of the European Communities Conf. on Photovoltaic Solar Energy, Luxembourg,
pp. 269–277.
Marti A. and Araujo G. (1996), ‘Limiting efficiencies for photovoltaic energy conversion
in multigap systems’, Solar Energy Mater. Solar Cells 43, 203–222.
Mathematica (2008), Version 7.0 Publisher: Wolfram Research, Inc. Champaign, Illinois.
Mihailetchi V. D., Koster L. J. A. and Blom W. M. (2004), ‘Effect of metal electrodes on the
performance of polymer:fullerene bulk heterojunction solar cells’, Appl. Phys. Lett.
85, 970–972.
Nozik A. J. (2008), ‘Fundamentals and applications of quantum-confined structures’, in
Archer M. D. and Nozik A. J. (eds), Nanostructured and Photoelectrochemical Systems
for Solar Photon Conversion, Imperial College Press, London.
Nozik A. J. (2010), ‘Nanoscience and nanostructures for photovoltaics and solar fuels’,
Nano Lett. 10, 2735–2741.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch02 page 85
86 M. A. Green
Trupke T., Green M. A. and Würfel P. (2002a), ‘Improving solar cell efficiencies by the
up-conversion of sub-band-gap light’, J. Appl. Phys. 92, 4117–4122.
Trupke T., Green M. A. and Würfel P. (2002b), ‘Improving solar cell efficiencies by down-
conversion of high-energy photons’, J. Appl. Phys. 92, 1668–1674.
Trupke T., Würfel P., Uhlendorf I. and Lauermann I. (1999), ‘Electroluminescence of the
dye-sensitized solar cell’, J. Phys. Chem. B 103, 1905–1910.
Trupke T., Zhao J., Wang A., Corkish R. and Green M. A. (2003), ‘Very efficient light
emission from bulk crystalline silicon’, Appl. Phys. Lett. 82, 2996–2998.
Vandewal K., Tvingstedt K., Gadisa A., Inganäs O. and Manca J. V. (2009), ‘On the origin of
the open-circuit voltage of polymer–fullerene solar cells’, Nature Mater. 8, 904–909.
Werner J. H., Brendel R. and Queisser H. J. (1995), ‘Radiative efficiency limit of terrestrial
solar cells with internal carrier multiplication’, Appl. Phys. Lett. 67, 1028–1030.
Whitehead A. N. (1911), An Introduction to Mathematics, Henry Holt & Co., New York,
Chapter V.
Wilson A. H. (1931), ‘The theory of electronic semiconductors’, Proc. Royal Soc. London
A133, 458–491.
Wolf M. (1960), ‘Limitations and possibilities for improvement of photovoltaic solar energy
converters’, Proc. IRE 48, 1246–1263.
Würfel P. (1982), ‘The chemical potential of radiation’, J. Phys. C 15, 3967–3985.
Würfel P. (1995), ‘Is an illuminated semiconductor far from thermodynamic equilibrium’,
Solar Energy Materials and Solar Cells 38, 23–28.
Würfel P. (2000), Olympic Conf. on Third Generation Photovoltaics, Sydney (unpublished
work).
Würfel P., Brown A. S., Humphrey T. E. and Green M. A. (2005), ‘Particle conservation in
hot-carrier solar cell’, Progr. Photovoltaics 13, 277–285.
Würfel P. and Wolfgang R. (1980), ‘Upper limit of thermophotovoltaic solar-energy
conversion’, IEEE Trans. Electron Devices ED-27, 745–750.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch03 page 87
CHAPTER 3
3.1 Overview
Front page headlines in the New York Times and the Wall Street Journal in 1954
heralded to the world the demonstration of the first reasonably efficient solar cells,
an announcement made possible by the rapid development of crystalline silicon
technology for miniaturised electronics.The majority of solar cells made since then
have been based on silicon in monocrystalline or large-grained polycrystalline
form. There are two main reasons for this. One is that silicon is an elemental
semiconductor with good stability and a well-balanced set of electronic, physical
and chemical properties, the same set of strengths that have made it the preferred
material for microelectronics. The second reason why silicon cells have been so
dominant is that the success of silicon in microelectronics created an enormous
industry where the economies of scale directly benefited the initially much smaller
photovoltaics industry.
In the early days of the industry, most silicon cells were made using thin
wafers cut from large cylindrical monocrystalline ingots prepared by the exacting
Czochralski (CZ) crystal growth process and doped to about one part per million
with boron during the ingot growth. A larger proportion now use what are referred
to as ‘multicrystalline’ wafers sliced from ingots prepared by a simpler directional
solidification technique, that produces large-grained polycrystalline ingots. To pro-
duce a cell, these boron-doped starting wafers generally have phosphorus diffused
at high temperatures a fraction of a micron into the surface to form the p-n junction
required for photovoltaic action. Metal contacts to both the n- and the p-type side
87
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch03 page 88
88 M. A. Green
of the junction are formed by screen printing metal pastes which are then densified
by firing at high temperature. Each cell is typically 12–20 cm either in diameter or
along either side if square or rectangular.
Cells generally are sold interconnected and encapsulated into a weatherproof,
glass-faced package known as a module, as shown in an exploded view in Fig. 3.1
Each module typically contains 72 cells soldered together in series. Each individual
cell gives a maximum output of about 0.63 V in sunlight. The output current depends
on cell size and the sunlight intensity (solar irradiance) but generally lies in the
4–12 A range in bright sunshine. The packaging consists of a glass/polymer lami-
nate with the positive and negative leads from the series-connected cells brought out
in a junction box attached to the module rear. Such modules have proved extremely
reliable in the field, with most manufacturers offering a 25-year warranty on the
module power output, one of the longest warranties provided for any commercial
product (saucepans have been suggested as one of the few manufactured products
rear cover
cells EVA
EVA
glass
Figure 3.1 Exploded view of a standard silicon photovoltaic module. The different layers
shown are laminated together under pressure at a temperature around 140–150 C where the
transparent EVA (ethylene vinyl acetate) softens and binds the different layers together on
cooling. Source: Green and Hansen (1998).
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch03 page 89
with a comparable warranty period — my most recently purchased pan only having
a 10-year warranty!).
The efficiency of the cells in the module would typically lie in the 15–19%
range, appreciably lower than the fundamental ‘detailed balance’ or Shockley–
Queisser limit of 33.5% for silicon (see Chapter 2). Module efficiency is slightly
lower than that of the constituent cells due to the area lost by frames and gaps
between cells, with module efficiency generally lying in the 12–15% range (Haase
and Podewils, 2011). Over the last few years, commercial cells and modules of
significantly higher performance have been available in multi-megawatt quantities
using more advanced cell processing technology, discussed in more detail later.
Cells of 20–22% efficiency and module efficiency in the 18–20% range are now
commercially available. (Unless otherwise noted, all efficiencies quoted in this
chapter are at standard test conditions, namely with a cell temperature of 25C
under 1000 W m−2 sunlight intensity with the standard global air mass 1.5 spectral
distribution.)
This chapter discusses the historical but steadily weakening links between
silicon solar cells and the broader microelectronics industry. Also discussed are
standard and improved methods for preparing silicon cell substrates and for
processing cells to extract as much performance as possible from these at the
lowest possible overall cost. The chapter also describes progress with supported
silicon films, which have the advantage of significantly reduced silicon material
requirements.
90 M. A. Green
p-type
(a)
(b)
n-type
p-type
(-)
(+) (+)
(c)
Figure 3.2 (a) Silicon solar cell reported in 1941 relying on ‘grown-in’ junctions formed
by impurity segregation in recrystallised silicon melts; (b) helium-ion bombarded junction
device of 1952; (c) first modern silicon cell, reported in 1954, fabricated on single-crystalline
silicon wafers with the p-n junction formed by dopant diffusion. Source: Green (1995).
92 M. A. Green
antireflection coating
top metal finger
n-type
(a)
thin
n-type
layer
p+ layer
p-type
(b)
textured surface
n-type
p-type
p+
rear contact
(c)
Figure 3.3 (a) Space silicon cell design developed in the early 1960s which became a
standard design for over a decade; (b) shallow junction ‘violet’ cell; (c) chemically textured
non-reflecting ‘black’ cell. Source: Green (1995).
a rapidly increasing number of satellites (Wolf, 1976). The basic cell design that
evolved (Fig. 3.3a), remained unchanged from the early 1960s for almost a decade.
In the early 1970s, a reassessment of cell design at COMSAT Laboratories
showed that a shallower diffusion combined with more closely spaced metal fingers
gave a substantial improvement in the cell performance by improving the response
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch03 page 93
to blue wavelengths (Lindmayer and Allison, 1973). The resulting cells, shown in
Fig. 3.3b, known as ‘violet’ cells due to their characteristic colour arising from the
shorter wavelengths reflected, produced the first improvement in cell performance
for over a decade.
This improvement was augmented by the realisation that incorporating a thin
heavily doped layer under the back contact, creating a so-called ‘back surface
field’, gave unexpected benefits (Godlewski et al., 1973). This approach worked
best if the rear doped layer was formed by alloying the underlying silicon with
aluminium deposited over the cell rear. Not long afterwards, the idea of using
anisotropic chemical etches to form geometrical features on the silicon surface was
successfully demonstrated, also at COMSAT Laboratories (Haynos et al., 1974),
and resulted in a further boost in performance, taking terrestrial cell efficiency to
above 17% (Fig. 3.3c). The surface features consisted of square-based pyramids
defined by slowly etching {111} crystallographic planes. These greatly reduced
reflection from the cell surface, giving these ‘black cells’ the appearance of black
velvet after antireflection coating.
The improvements of the early 1970s came about primarily by enhancing the
ability of the cell to collect carriers generated by the incoming photons. Since
cells now appeared to be performing to close to their full potential in this regard,
it seemed that any further improvement in cell performance would come from
increased open-circuit voltage. Gains in this area became the focus of work directed
at improving cell efficiency throughout the second part of the 1970s, largely as a
result of a program directed by NASA-Lewis aimed at targeting better space cell
performance (Brandhorst and Bernatowicz, 1980).
On the commercial front, the oil embargoes of the early 1970s generated
widespread interest in alternative sources of terrestrial energy. A small terrestrial
photovoltaic industry came into existence largely as a result of the US Govern-
ment’s photovoltaic program. One component of this program (Christensen, 1985),
arguably the most successful in terms of developing the industry and its products,
involved a staged series of purchases of photovoltaic modules meeting increasingly
stringent specifications.
The first such purchase in 1975/76, known as ‘Block I’, was remarkable for the
diversity of both cell fabrication and module encapsulation approaches used in the
product supplied by four different manufacturers. One manufacturer, Spectrolab
of Sylmar, California, supplied cells where the contacts had been applied using
screen printing (Ralph, 1975), the forerunner of the billions of cells of this type
which were to follow. In the ‘Block II’ purchases under this program (1976/77),
the same company combined screen-printed cells with a laminated module design
(Fig. 3.1), a combined approach that had been adopted by almost all commercial
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch03 page 94
94 M. A. Green
Figure 3.4 Screen-printed crystalline silicon solar cell (not to scale). Source: Green (1995).
manufacturers by the early 1980s and, with relatively minor modification, remains
the present commercial standard.
The main features of a commercial screen-printed cell are shown in Fig. 3.4.
The basic cell design is similar to that of a standard space cell of the 1960s
(Fig. 3.3a), but incorporates the surface texturing of the ‘black’ cell of Fig. 3.3c as
well as the screen-printing approach to applying the front and rear contacts. Present
efficiencies are quite similar to those demonstrated by ‘black’ cells in 1974.
Since the Block II purchases of 1976/77, no major changes were made in either
the basic screen-printed approach to cell fabrication or to the cell encapsulation
approach until quite recently. Considerable attention, however, has been directed
towards reducing the cost of the silicon wafer, initially grown by the Czochralski
technique, since this accounts for a large fraction of the cost of a standard silicon
module. The most successful approach has been the simplification of the ingot
growth processes by using cruder directional solidification approaches to produce
multicrystalline ingots (Ferrazza, 1995; Rodriguez et al., 2011).
The first multicrystalline silicon cells developed specifically for the terres-
trial market were reported in 1976 (Lindmayer, 1976; Fischer and Pschunder,
1976) and commercial multicrystalline cells have been available since the late
1970s. These multicrystalline approaches involve basically a reversion to the earlier
ingot-forming approaches for crystal rectifiers, techniques pre-dating the micro-
electronics explosion. In recent years, multicrystalline silicon cells accounted for
the majority of the total market for photovoltaic product (Hering, 2011). Another
major area of emphasis has been to reduce the thickness of the silicon wafer by
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch03 page 95
slicing it more thinly. This resulted in the replacement of inner diameter sawing
methods traditionally favoured by the microelectronics industry by wire-cutting
approaches, described in more detail in Section 3.3.1.
96 M. A. Green
finger
microgroove
n+ thin oxide
p
p+
back contact
(a)
oxide n+
n++
metal
(b)
Figure 3.5 (a) The microgrooved passivated emitter solar cell (PESC cell) of 1985, the first
silicon cell to exceed 20% efficiency; (b) buried contact solar cell. Source: Green (1995).
or other dielectric coating the top surface in this case not only serves as surface
passivation, but also as a diffusion mask to confine the heavy diffusion to the laser-
grooved areas and as a plating mask for the subsequent plating of metal into these
grooved areas. During the early 1990s, the buried contact approach produced the
highest performance terrestrial cells made in any appreciable volume, with effi-
ciency in the 17–18% range obtained using standard commercial silicon wafers
(Jordan and Nagle, 1994).
The next major laboratory improvement in silicon cell design came in the use
of oxide passivation along both the front and rear surfaces, first demonstrated in
the rear point-contact solar cell developed by Stanford University. As shown in
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch03 page 97
n+ busbar
p+ busbar
+
n
+
+ n
n
oxide
sunlight
Figure 3.6 Rear point-contact solar cell, which demonstrated 22% efficiency in 1988 (cell
rear shown uppermost). Source: Green (1995).
Fig. 3.6, this cell has an unusual design in that both positive and negative contacts
are on the rear surface of the cell. Although this might, at first sight, appear to
be a regression to a similar design to that used in the first modern silicon cell
of Fig. 3.2c, there is a substantial difference in the way the two types of cells
operate. The modern rear junction cells take advantage of the excellent quality of
silicon now available. Carrier diffusion lengths are several times the cell thickness,
allowing carriers photogenerated near the top surface of the cell to diffuse to the rear
contacts. In the earlier device, the junctions at top and rear surfaces are electrically
connected around the cell edge (Fig. 3.2c). Most carriers in this earlier cell are
collected by the top junction and flow around the cell edges to the rear contact. The
rear point-contact cell demonstrated 22% efficiency in 1988 and has since been
successfully commercialised, as subsequently discussed.
The next improvement in silicon cell efficiency came, again at UNSW, by
combining the earlier developments in the PESC cell sequence with the front and
rear oxide passivation first demonstrated in the rear point-contact cell. This is
possible in a number of ways as shown in Fig. 3.7. In the PERC cell (passivated
emitter and rear cell) of Fig. 3.7a, the first successfully demonstrated, rear contact
is made directly to the silicon substrate through holes in the rear passivating oxide.
This approach works reasonably well provided the substrate is sufficiently heavily
doped to ensure low contact resistance between the metal and substrate (below
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch03 page 98
98 M. A. Green
n+ n oxide
(a)
p-silicon
p+
n+ n oxide
(b) p-silicon p+
p+ p+
+
p
n+ n oxide
(c) p-silicon p
+
p+ p+ p
p+
n+ p-silicon
oxide
n p+
(d) p+ p+
Figure 3.7 A family of four related high-efficiency solar cell structures: (a) the passivated
emitter and rear cell (PERC cell); (b) the passivated emitter, rear locally diffused cell (PERL
cell) which took efficiency above 24% in the early 1990s and subsequently to 25%; (c) the
passivated emitter, rear totally diffused cell (PERT cell); and (d) the passivated emitter, rear
floating junction cell (PERF cell). Source: Green and Hansen (1998).
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch03 page 99
about 0.5 cm resistivity for p-type substrates). The PERC cell is often suggested
as a relatively low cost way for making silicon cells above 20% efficiency, since it
is the simplest of the approaches of Fig. 3.7.
Historically, the next improvement was demonstrated by the PERL cell (pas-
sivated emitter, rear locally diffused) of Fig. 3.7b. In this case, local diffusion is
used in the area of the rear point contact to provide a minority-carrier-reflecting
region between this contact and the substrate and to reduce contact resistance. This
approach produced the first 24% efficient silicon cell in 1994 (Zhao et al., 1995) and
the world best result for silicon of 25% (Green, 2009) under the standard AM1.5
global spectrum. The PERL cell has been used in reasonably large quantities in
solar car racing and in space cells.
The third cell of Fig. 3.7c is the PERT cell (passivated emitter, rear totally
diffused). This has given performance almost equivalent to the PERL cell and also
offers some fabrication simplifications. The PERT cell has also been used in space
cell production. The final structure shown in Fig. 3.7d, the PERF cell (passivated
emitter rear floating junction) offers perhaps the best long-term potential for high
performance. This structure has produced the highest open-circuit voltage of the
cells of Fig. 3.7, with open-circuit voltage up to 720 mV demonstrated under stan-
dard test conditions (Wenham et al., 1994), together with efficiencies above 23%.
One feature of these cell designs is the very effective trapping of light within
the cell. By depositing metal over the entire rear surface of the cell but ensuring it
is displaced from the silicon substrate by an intervening layer of oxide, very high
rear reflectance is obtained for light striking this rear reflector from within the cell.
When combined with appropriate geometrical structure on the front surface of the
cell, weakly absorbed light that is reflected from this rear surface can be trapped
quite effectively within the cell, taking advantage of total internal reflection from the
front surface. This greatly extends the response of the cell to infrared wavelengths.
Cells that convert infrared wavelengths with an efficiency approaching 50% have
been demonstrated (Green et al., 1992).
100 M. A. Green
and nitride passivation of phosphorus-diffused surfaces also give good results. This
makes phosphorus the preferred dopant for the more critical top doped layer of the
cell. For the rear surface, there are fewer constraints on the required properties of
the diffusion, allowing boron diffusions to work well there.
The Stanford University rear point-contact cell of Fig. 3.6 is shown as origi-
nally conceived, without any diffusion on the surface exposed to sunlight. Adding a
non-contacted or ‘floating’ phosphorus diffusion to the top (sunlight-exposed) sur-
face improved performance, with 22.7% efficiency soon thereafter demonstrated,
long the best for any cell on an n-type silicon substrate.
In attempting to adapt their p-type wafer work to n-type substrates, the UNSW
group were able to equal this result, but again using a phosphorus diffusion along
the top surface. A ‘front surface field’ structure was used, as in the PERT cell of
Fig. 3.7c, but with the substrate n-type rather than p-type. The standard PERL
structure of Fig. 3.7b on n-type substrates with all polarities reversed gave slightly
poorer results due to the more difficult challenge involved in passivating the B-
doped top surface.
The challenge of improving cell performance on n-type substrates was sub-
sequently met by the combined efforts of the Fraunhofer Institute in Freiburg,
Germany, and ECN, Netherlands (Benick et al., 2008; 2009). Dielectrics such as
silicon oxide and nitride work well on n-type surfaces since they tend to have nett
positive electrical charge within them, like many other dielectrics, with this charge
tending to attract electrons to underlying silicon surfaces (Godfrey and Green,
1980). This also depletes holes in the surface region, reducing surface recombi-
nation rates by squeezing off the supply of these minority carriers. It was found
that Al2 O3 could be prepared with nett negative charge, repelling electrons from
the interface to the interior, similarly restricting the supply for recombination on
p-type surfaces, such as for the top surface of an n-type cell if B-doped. A sig-
nificant improvement in n-type cell performance to 23.4% was demonstrated by
using the PERL cell structure of Fig. 3.7b with all polarities reversed but with the
top surface coating (normally silicon oxide under a ZnS/MgF double layer antire-
flection coating) replaced by 30 nm Al2 O3 overlaid by 40 nm Si3 N4 (Benick et al.,
2009). Although below the best PERL cell on p-type substrates on each of the
major cell parameters (Voc , i sc , FF), this result is important in demonstrating the
ability to passivate B-doped surfaces, increasing future flexibility in cell design.
Another significant development has been the HIT (heterojunction with intrin-
sic thin layer) cell, developed by Sanyo (now Panasonic) and shown in Fig. 3.8
(Tanaka et al., 1993; Sawada et al., 1994; Maruyama et al., 2006). Here the
problems raised by diffusions and their passivation are avoided totally by using
deposited layers of hydrogenated amorphous silicon (a-Si:H) for both positive
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch03 page 101
metal
TCO
+
p (a-Si)
i(a-Si)
n(c-Si)
i(a-Si)
+
n (a-Si)
Figure 3.8 HIT (heterojunction with intrinsic thin layer) cell on textured crystalline silicon
substrate (after Green and Hansen, 1998).
and negative contacts to top and rear surfaces. The hydrogen, incorporated in the
5–10% range by atomic volume, modifies the electronic structure of the a-Si:H
alloy, producing much larger bandgap (about 1.7 eV compared to 1.1 eV for crys-
talline silicon). The a-Si:H can be doped n- or p-type, although the electronic
quality of doped material tends to be poor. This different form of silicon is widely
used in its own right for thin-film solar cells (Chapter 4), making the HIT cell
structure a marriage between two successful cell technologies. Features such as
transparent conducting oxides (TCO) are also used, with these widely used in thin-
film technologies but not in standard silicon technology. The HIT also has some
similarities with III–V cell technology (Chapter 7), where heterojunctions between
semiconductors with different bandgaps are a common feature.
The advantage of using wider bandgap material is that minority-carrier con-
centrations are greatly suppressed by the wider bandgap, correspondingly reducing
recombination rates. In the HIT cell structure, all significant recombination occurs
in the wafer substrate or at its surfaces. An important additional feature of the HIT
cell is a very thin layer of intrinsic a-Si:H, only 10–20 nm wide, interposed between
the wafer and the thin doped a-Si:H regions. This thin intrinsic layer helps reduce
interfacial recombination, possibly by reducing defect states in the bandgap of the
a-Si:H in these regions that would be created by doping.
The HIT cell has a number of attractive features. One is that all processing of the
starting wafer can be effected at low temperature, allowing the initial wafer quality
to be retained during processing. The resulting cells have a bifacial structure, so
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch03 page 102
102 M. A. Green
can respond to light from either direction and, in principle, wafers of either polarity
could be used, although n-type wafers are the preferred choice.
Standard HIT cells have also demonstrated outstanding efficiencies, up to
24.7% for large area cells (Maruyama et al., 2006), until recently the best ever
result for n-type substrates. The outstanding performance parameter has been the
open-circuit voltage, with values up to 745 mV demonstrated for 23% efficient cells
(compared to 708 mV for 25% efficient PERL cells). This high voltage results from
inherently lower surface recombination due to the passivation effects of the overly-
ing a-Si:H. Since there is a very direct and general relationship between the open-
circuit voltage and the temperature coefficient of performance of the cell (Green,
2003), this gives a significant advantage at high temperature compared with stan-
dard commercial silicon cells (where Voc might be only in the 610–630 mV range).
Technically, the main disadvantages with the standard HIT cell approach are
optical. The top doped a-Si:H layer absorbs ultraviolet and blue light and, because
of its poor electronic quality, this absorption does not contribute to the cell current.
Similarly, the conducting oxide layer will absorb incoming photons, particularly at
the infrared end of the spectrum. This will additionally degrade the effectiveness
of any light-trapping scheme that attempts to boost the cell’s response to poorly
absorbed infrared light. Consequently, current output from the cell is 5–10% lower
than from other similarly high-efficiency cells. The cost of processing wafers to
cells is also reported to be moderately high (Song et al., 2010). The HIT approach
has been important in demonstrating that almost perfect surface passivation is
possible in silicon cells, showing what ultimately might be possible in the future.
The final recent improvement to be discussed relates to the Stanford rear
point-contact cell (Fig. 3.6), an important contributor to the evolution of sili-
con cell efficiency, as already noted. This cell was commercialised by SunPower
(De Ceuster et al., 2007) and now produces the highest efficiency commercial
panels. A recent re-evaluation of cell design produced impressive results (Cousins
et al., 2010), also for n-type substrates. Figure 3.9 shows the cell structure of this
‘Generation III’ product.
Improved diffusions for both contacted and non-contacted regions of the rear
of the cell have contributed to the improved performance (Cousins et al., 2010).
A combination of oxide/nitride passivation is used for non-contacted areas where
diffusions are quite light (∼1018/cm3 ). For the contact regions, a heavier doping is
used as a compromise between the optimum for those regions directly contacted by
metal and those, also in the contact region, but not so directly contacted. An impres-
sive efficiency of 24.2% was demonstrated on a large-area, commercially-sized
cell (156 cm2 ) using commercial production equipment, with an impressive Voc of
721 mV (Cousins et al., 2010).
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch03 page 103
Figure 3.9 SunPower ‘Generation III’ cell. Source: Green and Ho (2011).
30
lambertian
Efficiency (%)
20
planar, one pass
10
0
1 10 100 1000 10000
Thickness (µ m)
Figure 3.10 Limiting efficiency of a silicon solar cell as a function of cell thickness with
and without Lambertian light trapping (global AM1.5 spectrum, 100 mW cm−2 , 25◦ C).
Source: Green (1995).
104 M. A. Green
become more severe. Interestingly, Fig. 3.10 shows that the commercial imperative
to decrease wafer thickness to decrease silicon costs (SEMI, 2011) may actually
lead to higher energy conversion efficiency, with appropriate device structures.
Various approaches have been suggested that have the potential, in principle,
for exceeding even the challenging efficiency limits of Fig. 3.10. These include
the use of tandem cells, the use of high-energy photons to create more than one
electron-hole pair (Werner et al., 1994), or the use of sub-bandgap photons in
schemes such as incorporation of regions of lower bandgap (Healy and Green,
1992), multiple quantum wells (Barnham and Duggan, 1990), mid-gap impurity
levels (Wolf, 1960) or up- and/or down-conversion (Trupke et al., 2002a; 2002b).
The tandem cell approach appears the most likely to have impact in the long term,
once the problems are overcome with lattice-matching a top cell with a suitable
bandgap to silicon.
106 M. A. Green
diameter has increased to the required size, the process is stopped and the rods
mechanically broken into smaller chunks, that maintain ‘nine-nines’ purity. These
chunks then become the starting point for the growth of ingots of good crystalline
quality.
In the 1970s and early 1980s, several other options for preparing silicon feed-
stock were investigated as part of a large US government photovoltaic program
supported by the Carter administration (Christensen, 1985). Approaches investi-
gated ranged from those involving radically different techniques to those exploring
only minor changes from the sequence outlined above, such as the use of different
compounds of silicon as the intermediate during the purification process. One such
process, based on the use of silane as the intermediate (Christensen, 1985), is now
used commercially. Decomposition of the intermediary gas in fluidised bed reac-
tor was also studied in this program and is now also used commercially, greatly
reducing process energy requirements. Other projects were based on the use of
upgraded metallurgical grade (UMG) silicon, a line of investigation still attracting
great attention given its appeal in circumventing the complexity of the traditional
purification approach.
seed
crystal
(a)
sawn crystal
(recycled)
wafers
(b)
Figure 3.11 (a) Czochralski (CZ) growth; (b) squared-off CZ ingot. Source: Green and
Hansen (1998).
around their outer perimeter. The cutting surface is a diamond impregnated edge
surrounding a hole within the tensioned metal sheet. This technique gives excellent
dimensional tolerance, although there are limitations arising from the thickness of
the silicon wafers that is possible to produce while still maintaining high yield.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch03 page 108
108 M. A. Green
Inner-diameter
saw blade
ingot
to spool
wafer
(a)
ingot
guide
(b)
Figure 3.12 (a) Inner diameter wafer sawing; (b) continuous wire sawing. After Dietl et al.,
1981.
Other limitations arise from the wastage of silicon as ‘kerf’ loss during cutting.
Generally, about 10–15 wafers per centimetre of ingot length were achieved by
this process.
An alternative technique now almost universally used in photovoltaics is based
on wire sawing (Fig. 3.12b). In this case, tensioned wire is used to guide an abrasive
slurry through the ingot (Schumann et al., 2009). Advantages are thinner wafers
and less surface damage for these wafers as well as lower kerf or cutting loss,
initially allowing the sawing of over 20 wafers per centimetre. With subsequent
reduction in wafer thickness to 160–180 microns in 2011 and of the diameter of
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch03 page 109
the wire to the 100–120 micron range, this figure has increased to over 30 wafers
per centimetre, with up to 50 wafers per centimetre of ingot length expected over
the coming decade.
There has been a recent resurgence of interest in an alternative wire sawing
approach where the wire is similarly thin, but impregnated with diamond particles
(Schmidt, 2011). This approach has the advantage of much higher cutting speed as
well as reduced kerf loss. The process also offers more potential for the recycling of
the silicon that would otherwise be wasted as kerf loss during the sawing process.
An alternative to the standard Czochralski process for producing crystalline
ingots is the floatzone (FZ) process. Although some studies have predicted superior
economics when compared to the Czochralski process for cell production due
to the elimination of consumables such as quartz crucibles, the FZ process, as
commercially implemented, is capable of accepting feedstocks only in the form of
high-quality cylindrical rods. This makes it unsuitable for using low-cost source
material.
Early interest in producing monocrystalline silicon by a directional solidifi-
cation process (Schmidt, 2011) has been recently revived and will be discussed
in the next section. Unlike the FZ process, this approach is very tolerant of low-
grade source material (Khattak et al., 1981), similar to the directional solidification
approaches to producing multicrystalline silicon, also to be discussed in the next
section.
110 M. A. Green
silicon
mould
(a)
(b)
Figure 3.13 (a) Directional solidification of silicon within a mould; (b) sawing of large
ingot into smaller sub-sections. Source: Green and Hansen (1998).
a tonne (Rodriguez et al., 2011). This weight is expected to double over the coming
decade (SEMI, 2011).
The large ingots are sawn into smaller sections as shown in Fig. 3.13b, which
eventually give wafers generally 15.6–20 cm along the sides. These smaller sec-
tions can be sawn by the standard continuous wire sawing processes. The resulting
multicrystalline wafers are capable of producing cells of over 85% of the perfor-
mance of a monocrystalline cell fabricated on a CZ wafer. However, because of
the higher packing density possible due to their square or rectangular geometry,
this performance difference is largely masked at the module level, with multicrys-
talline module performance lying in the range demonstrated by modules made
from monocrystalline cells (Haase and Podewils, 2011).
Interest in producing monocrystalline rather than multicrystalline silicon
ingots by such directional solidification approaches dates back to the 1970s
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch03 page 111
112 M. A. Green
Figure 3.14 (a) Edge-defined, film-fed growth (EFG) method; (b) growth of a nonagonal
ribbon of silicon using the EFG method. Source: Green and Hansen (1998).
from this material have been available sporadically since the early 1980s. However,
after building up to a manufacturing capacity of about 200 MW/year, a decision
was made to cease production in 2009 since ‘ingot technology has now proven to
be a more economical process’ (Schott, 2009).
An even older ribbon growth process is the dendritic web approach of Fig. 3.15,
first described by Westinghouse in the 1960s. In this approach, close thermal control
is used to cause two dendrites spaced several centimetres from each other to solidify
first during the growth step. When these are drawn from the melt, a thin sheet of
molten silicon is trapped between them. This quickly solidifies to form a ribbon.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch03 page 113
silicon dendrites
or carbon
string
molten
silicon
Figure 3.15 Schematic illustrating the dendritic web growth process and the string ribbon
approach. Source: Green and Ho (2011).
A somewhat related approach is the string ribbon approach. In this case, the
molten silicon is trapped between two graphite strings that are drawn from the
melt. This relaxes the requirement on thermal control, compared with the previous
dendritic web approach. The string ribbon approach has been commercialised by
Evergreen Solar (Janoch et al., 1997; Wallace et al., 1997; Rodriguez et al., 2011).
Evergreen produced an estimated 157 MW of modules using this technology during
2010 while a formerly related company, Sovello, produced an additional 145 MW
(Hering, 2011).
Despite the obvious attractions of the ribbon approaches, market share steadily
declined, representing only 1.2% of production in 2010 after peaking at 5.6% in
2001 (Hering, 2011), with production now phased out. This highlights the difficulty
of establishing unique technology in an industry where mainstream manufacturers
have access to a skill and resource base far larger than that of any individual
manufacturer.
114 M. A. Green
02 + N2
Silicon wafers
Carrier
gas
Figure 3.16 Phosphorus diffusions process. Source: Green and Hansen (1998).
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch03 page 115
To break the connection between the phosphorus diffused into front and rear
surfaces, an ‘edge junction isolation’ step is needed to remove the thin phosphorus
layer around the edge of the wafer. This isolation was once achieved by ‘coin
stacking’ the wafers so that only their edges were exposed. The stack was then
placed in a plasma etcher to remove a small section of silicon from the wafer edge,
hence breaking the conductive link between front and rear surfaces. Single-sided
etching is now the preferred approach in which the wafer sits on the etching solution
so that only the desired surface is etched.
A silicon nitride quarter-wave antireflection coating is applied to the cell at
this stage. Not only does silicon nitride have a reasonably good refractive index
for this application, it is deposited using silane (SiH4 ) and consequently has a high
hydrogen content. This hydrogen can be released into the silicon substrate, giving
a beneficial defect passivating effect, particularly for multicrystalline cells.
Cell processing is completed by the screen printing of metal contacts onto the
front and rear surface. Silver and aluminium paste (consisting of a suspension of
fine metal particles and glass frit in an organic medium together with appropriate
binders) is squeezed through a patterned screening mesh onto the cell surface. After
application, the paste is dried at low temperature. The Ag-based paste for the top
surface is printed in a characteristic finger pattern shown in Fig. 3.17 to minimise
the resistive losses in the cell while allowing as much light as possible into it. For
the rear, two different pastes are generally applied as shown. An Ag-rich paste is
applied to a small fraction of the rear to allow soldering of the cell interconnections,
while most of the rear is covered by an Al-based paste. After these three layers
front
B
rear
Figure 3.17 Standard 156 mm × 156 mm solar cell ‘H’ metallisation pattern. Screen-printed
silver regions are shown coloured black. The pattern consists of six unit cells of size A × B.
Both front and rear views are shown. Source: Green and Ho (2011).
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch03 page 116
116 M. A. Green
have been individually screened and dried, the pastes are then fired at a higher
temperature to drive off the remaining organics and to allow the metal particles
in the paste to coalesce. The glass frit is important in promoting adhesion to the
silicon substrate. Often pastes are doped with dopants such as phosphorus, in this
case to help prevent the screened contact from penetrating the thin phosphorus skin
that it is intended to contact.
This screen-printing method for applying the metal contact was borrowed
in the early 1970s from the hybrid microelectronics industry (Ralph, 1975). This
ensured the ready availability of both screen-printing equipment and the paste firing
furnaces suited to this application. Labour and equipment costs associated with this
step tend to be very low. However, the pastes themselves can be expensive (Green,
2011) and an even larger cost penalty is paid for the simplicity of this approach by
the forfeiture of the inherently available power output from the silicon wafer, as
discussed later.
After the paste firing, the cells are then ready for testing under a solar simulator.
Cells are usually graded based on their short-circuit current or current at a nominal
operating voltage, e.g. 500 mV. Generally, cells are sorted into 5% performance
bins. This sorting is required to reduce the amount of mismatch between cells
within the completed module. To a large extent, the output current of the module is
determined by that of the worst cell in the module, resulting in large power losses
within mismatched modules. Even worse, low-output cells can become reverse-
biased under some modes of module operation and destroy the module by localised
over-heating.
Silicon ribbon substrates often require modifications of the above standard
sequence. For example, the rippled surface which is a natural consequence of
the EFG ribbon growth process poses continuity hazards for screen-printed met-
allisation. To accommodate this rough surface, a novel technique was developed
whereby the metal paste was squeezed through an orifice and then dropped to the
cell surface, much the same as squeezing toothpaste from its tube onto a toothbrush.
118 M. A. Green
120
PV17A
19 PV15A 110
18 fire- 90
through PV173
80
TiO2
17 70
SiO2 ohms/sq.
AR 60
coat
16 50
40
15 30
1997 2000 2003 2006 2009 2012 2015
Figure 3.18 Past and planned product introduction schedule for one major paste supplier.
Shown are the different generations of paste and the emitter sheet resistivity the pastes are
designed to accommodate. The author’s estimate of likely production efficiency is based
on unscaled estimates provided by the paste manufacturer. Adapted from Laudisio et al.
(2010).
finger layer 2
layer 1
n-type n-type
n+
p-type p-type
p+ p+
rear contact
(a) (b)
Figure 3.19 (a) ‘Selective emitter’ approach; (b) ‘double printing’ to increase aspect ratio
of metal fingers. Source: Green and Ho (2011).
silicon can be optimised for this role while the second layer can be optimised for
good conductivity.
Recent rapid increases in silver prices cast doubt on the long-term role of this
metal in photovoltaics (Green, 2011). The industry seems likely to transition to
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch03 page 119
the use of Cu as the main conductive metal in the cell. Plating Cu to thin screen-
printed Ag lines is one possible approach, or it may be that the need for change
will stimulate the adoption of more innovative technology.
n+
n++ p
p+
metal
Figure 3.20 Semiconductor finger solar cell. Heavily doped semiconductor regions provide
an additional finger layer that is contacted by the traditional metal finger layer. Source: Green
and Ho (2011).
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch03 page 120
120 M. A. Green
Figure 3.21 Laser-doped selective emitter (LDSE) solar cell. Metal is selectively plated
to the heavily doped regions exposed during the laser doping step. Source: Green and Ho
(2011).
for top finger contact by removing the dielectric coating by melting the underlying
silicon, while simultaneously introducing dopants into the molten region. The front
contacts can then be plated as a Ni/Cu/Ag tri-layer. The cell has the advantages
of a lightly doped top surface region and of fine, highly conductive metal fingers.
Efficiency in the 19–20% range has been confirmed for cells on large area CZ
wafers fabricated using largely commercial equipment. This is expected to improve
to the 21–22% range once a similar contacting approach is applied to the rear of
the cell.
p+
n+
p+ p+
(a)
p+
n+
p+ p+
(b)
Figure 3.22 (a) Emitter wrap-through (EWT) solar cell; (b) metal wrap-through (MWT)
cell. Source: Green and Ho (2011).
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch03 page 122
122 M. A. Green
production. This was the first commercial cell of efficiency above 20%. In 2008,
SunPower introduced an improved ‘Generation II’ product which took efficiency
to 22% (De Ceuster et al., 2007). This has produced commercial modules in large
volume with 18.4–19.7% total area efficiency, the highest on the market (Haase
and Podewils, 2011). More recently, a Generation III product has been reported
(Cousins, 2010) with 24% cell efficiency that is likely to take commercial module
efficiency beyond 20% for the first time. Cell processing costs are reported to be
higher than for the standard screen-printing approach (Song et al., 2010).
Commercialisation of EWT cells is not as far developed as the rear-junction
cell or even the MWT device. A key disadvantage of both compared to the rear-
junction device is the need to drill multiple holes through the wafer using a laser,
although fewer are required for MWT devices. However, both EWT and MWT
cells are more tolerant of low-quality wafers than rear-junction devices. In fact,
the first solar modules of over 17% efficiency employing multicrystalline wafers
used MWT cells (Lamers et al., 2011). In 2010, at least two manufacturers offered
MWT devices commercially (Solland, 2009; Photovoltech, 2010), although man-
ufacturing volume was not large.
least, not published), but only seven selected combinations, of which Table 3.1
shows five.
From this table, several key results can be deduced. When comparing screen-
printed cells on ribbon (EFG), multicrystalline (DS) and monocrystalline (CZ)
wafers, the ribbon produced the lowest cost of 0.71/Wp followed by the mul-
ticrystalline wafers at 0.91/Wp and the monocrystalline wafers at 1.25/Wp. The
advantage of the ribbon stemmed almost entirely from the fact that it does not need
to be sawn, as previously mentioned. As noted earlier, EFG production ceased in
2009 due to unfavourable economics. This is probably as a result of the relatively
small effort that could be devoted to the development of this technology by a single
company compared with the huge effort devoted by almost the entire industry to
the directional solidification approach.
Comparing the different processing approaches on single-crystal wafers, the
cheapest is the buried contact at 1.15/Wp, followed by the screen-printed at
1.25/Wp, followed by the PERL at 1.78/Wp. The buried contact achieved its
cost advantage over the screen-printing approach by virtue of the increased effi-
ciency giving more power per unit processing area. In terms of module costs, the
standard laminated module approach was calculated to be slightly cheaper than an
alternative resin-fill approach.
At the time these results were published, many regarded the costs deduced
as impossibly ambitious. However, actual developments have shown them to be
conservative. In 2011, average module manufacturing cost across the industry was
estimated as US$1.56/Wp (Bolman et al., 2011), or about 1.08/Wp, with the
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch03 page 124
124 M. A. Green
The trend towards thinner cells, which arises mainly from efforts to reduce the
costs of the silicon wafer, may actually help to improve the cell efficiency, as already
noted. Thin wafers give the opportunity for back-surface fields or other rear-surface
passivation approaches to be used to improve cell performance, primarily through
increased voltage output. Given better feedstock material or cells below 150 µm
thickness, improved rear-surface passivation approaches such as demonstrated by
the PERC and PERL cells of Fig. 3.7, as well as rear-junction and HIT cells, will
become increasingly relevant.
126 M. A. Green
30
no limit
680 mV
25
Efficiency (%)
720 mV
20 600 mV
640 mV
15
1 10 100 1000 10000
Thickness ( µ m)
Figure 3.23 Limiting efficiency of silicon cell with Lambertian light trapping as a function
of surface recombination velocity, characterised in terms of the voltage limit imposed by
this recombination. Source: Green (1999).
50.5%
47.5%
42.5% 45%
40
AM1.5G Efficiency
33%
30
29%
20
10
Free choice 0
0 1 2 3
Si bottom cell
Number of cells
Figure 3.24 Efficiency limits for cells and cell stacks with silicon as the lowermost cell
compared to the unconstrained limit for series-connected cells. While silicon is a good choice
for a single-cell material with an efficiency limit of 29% compared to the unconstrained limit
of 33%, it is an even better choice for the lowermost cell in 2- and 3-cell stacks. Source:
Green and Ho (2011).
The tandem cell approach perhaps remains the most likely option, pending
technical improvements that allow III–V or related cell technologies to be grown
on silicon. Silicon has an ideal bandgap to be the lowermost cell in a 2–4 cell
tandem (Fig. 3.24) and also can now be prepared very inexpensively in wafer
form. It would provide a clean, low-cost substrate for the subsequent growth of
high-performance cells. Low-cost deposition of the subsequent cell layers would
be essential. Progress has been reported in depositing lattice-matched III–V com-
pounds on silicon (Grassman et al., 2009), as well as fully-relaxed Ge layers on
silicon that might serve as templates for the subsequent growth of a wider range
of III–V materials (Wistey et al., 2007; Tsao, 2011).
128 M. A. Green
substrate (Fang et al., 1974). A surprisingly large grain size was obtained, attributed
to eutectic reaction with the aluminium. In more recent times, laser crystallisa-
tion has been used in the active matrix liquid crystal display industry to produce
relatively small-grain polycrystalline silicon films from amorphous silicon pre-
cursors, generally deposited by low-pressure chemical vapour deposition. Grain
sizes are typically less than a micron or so, so that these films would probably
not be suitable for photovoltaics. Also, thicknesses for the active matrix display
industry tend to be only about 100 nm, which would be too thin for photovoltaic
application.
From 1989, a group at Sanyo explored the use of low-temperature solid phase
crystallisation of amorphous silicon as a technique for producing thin-film poly-
crystalline silicon cells. Good results were obtained, with 9.2% (unconfirmed) effi-
ciency reported in 1995 (Baba et al., 1995). These cells were approximately 1 cm2
in area deposited onto a textured metallic substrate and heated at approximately
600◦ C for many hours to enable the crystallisation of the originally amorphous
films. After crystallisation, the HIT structure developed by Sanyo (Fig. 3.8) was
used to complete the cell processing at low temperature.
A similar approach has been developed for depositing thin polycrystalline
films of silicon on glass (CSG), as described in more detail elsewhere (Green
et al., 2004). This produced small area modules of 10–11% efficiency (Keevers
et al., 2007). About 10 MW of 1.4 m2 modules have been produced commercially
with this approach.
Two groups pioneered the development of silicon thin films deposited directly
in microcrystalline form (µc-Si) onto glass substrates. In 1997, the University of
Neuchatel reported efficiencies of about 7% for 3 µm thick microcrystalline cells
deposited at 500◦ C (Shah et al., 1997). The cell had a p-i -n structure with the
intrinsic region comprising most of the device thickness. The cell was designed
for the intrinsic region to be depleted during normal device operation to create a
high electric field to aid carrier collection, as with a standard amorphous silicon
cell. Kaneka Corporation (Yamamoto et al., 1997) reported efficiencies of over
10% with a similar device structure (Fig. 3.25). Nearly the same efficiency has
been obtained with the total device thickness varied over the 1.5–3.5 µm range.
Both the above groups reported even higher efficiencies when amorphous sil-
icon cells were used in a tandem configuration on top of the microcrystalline
device (Fischer et al., 1997). Since this early work, the latter tandem device has
been commercialised by many groups with tandem a-Si:µc-Si cell modules of
typically 8–9% efficiency available from several sources (Haase and Podewils,
2011).
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch03 page 130
130 M. A. Green
Ag
ITO
p µ c-Si
i poly-Si
n poly-Si
electrode
glass
Figure 3.25 Structure of 10.1% efficient thin-film microcrystalline solar cell developed by
Kaneka. Cell thickness is typically 1–3 µm. After Yamamoto et al. (1997).
3.8 Summary
Although crystalline silicon devices have dominated the commercial marketplace
for photovoltaics for more than three decades, there still remains scope for consider-
able improvement in both the performance and cost of these cells. Current expec-
tations are that manufacturing costs below US$0.50/Wp will soon be achieved,
without major changes in present processing sequences. The trend towards thin-
ner silicon wafers to decrease wafer cost is compatible with ongoing increases
in cell efficiency provided new cell structures are adopted and improved methods
are demonstrated in production for passivating the rear surface of the cell. The
large potential for both performance and cost reduction will make these bulk sili-
con approaches an increasingly challenging target for thin-film approaches. In this
context, good progress continues to be made with supported silicon film.
Ultimately, some way of boosting the performance of silicon cells would seem
to be required for these to continue to maintain their dominance over the coming
decades. The most promising at present would seem to be a tandem cell stack.
Such structures would take advantage of the enormous recent cost reductions in
producing silicon wafers, which could be used as clean substrates for the deposition
of thin, high-quality layers of strongly absorbing, wider-bandgap material. Such
an approach would combine the robust commercial infrastructure of silicon cell
manufacturing with the outstanding laboratory progress with tandem cell stacks.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch03 page 131
Acknowledgement
This work has been supported by the Australian Government through the Australian
Renewable Energy Agency (ARENA). Responsibility for the views, information
or advice expressed herein is not accepted by the Australian Government.
References
AMG (2010), ‘AMG acquires solar silicon casting technology from BP Solar International
Inc.’, Press Release, 22 December.
Authier B. (1978), ‘Poly-crystalline silicon with columnar structure’, Festkörper-probleme
XVIII, Freidr. Vieweg & Son, Wiesbaden.
Baba T., Shima M., Matsuyama T., Tsuge S., Wakisaka K. and Tsuda S. (1995), ‘9.2%
efficiency thin-film polycrystalline solar cell by a novel solid phase crystallisation
method’, Conf. Record, 13th European Photovoltaic Solar Energy Conf., Nice, October,
pp. 1708–1711.
Barnett A. M., Ford D. H., Checchi J. C., Culik S., Hall R. B., Jackson E. L., Kendall C. L.
and Rand J. A. (1997), ‘Very-large-area silicon-filmTM solar cells’, Conf. Proc., 14th
European Photovoltaic Solar Energy Conf., Barcelona, pp. 999–1002.
Barnett A. M., Hall R. B., Fardig D. A. and Culik J. S. (1985), ‘Silicon-film solar cells on
steel substrates’, Conf. Record, 18th IEEE Photovoltaic Specialists Conf., Las Vegas,
October, pp. 1094–1099.
Barnham K. W. J. and Duggan G. (1990), ‘A new approach to high-efficiency multi-band-gap
solar cells’, J. Appl. Phys. 67, 3490–3493.
Benick J., Hoex B., Dingemans G., Richter A., Hermle M. and Glunz S. (2009), ‘High
efficiency n-type silicon solar cells with front side boron emitter’, Proc. 24th European
PVSEC, September, Hamburg, 863–870.
Benick J., Hoex B., van de Sanden M. C. M., Kessels W. M. M., Schultz O. and Glunz S.
W. (2008), ‘High efficiency n-type Si solar cells on Al2 O3 -passivated boron emitters’,
Appl. Phys. Lett. 92, 253504.
Bergin D. O. (1980), ‘Shaped crystal growth — a selected bibliography’, J. Crystal Growth
50, 381–396.
Bolman C., Coffey V., Fu B., Song J., Trangucci R. and Zuboff G. (2011), ‘The true cost of
solar power: the pressure’s on’, Photon Consulting, www.photonconsulting.com.
BP Solar (1991), Data Sheet, BP Saturn Solar Cells.
Brandhorst H. W. and Bernatowicz D. T. (1980), ‘Space solar cells: high efficiency and
radiation damage’,Conf. Record, 14th IEEE Photovoltaic Specialists Conf., San Diego,
pp. 667–671.
Bruton T. M. (2002), ‘MUSIC FM five years on: fantasy or reality?’, Paper OA6.1, Proc.
PV in Europe, October, Rome.
Bruton T. M., Luthardt G., Rasch K.-D., Roy K., Dorrity I. A., Garrard B., Teale L., Alonso
J., Ugalde U., Declerq K., Nijs J., Szlufcik J., Rfauber A., Wettling W. and Vallera A.
(1997), ‘A study of the manufacture at 500 MWp p.a. of crystalline silicon photovoltaic
modules’, Conf. Record, 14th European Photovoltaic Solar Energy Conf., Barcelona,
pp. 11–16.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch03 page 132
132 M. A. Green
Chapin D. M., Fuller C. S. and Pearson G. O. (1954), ‘A new silicon p-n junction photocell
for converting solar radiation into electrical power’, J. Appl. Phys. 8, 676–677.
Christensen E. (1985, ed.), Final Report, Flat Plate Solar Array Project: 10 Years of
Progress, Jet Propulsion Laboratory, Pasadena, CA, JPL Publ. 400–279.
Chu T. L. (1977), ‘Silicon films on foreign substrates for solar cells’, J. Crystal Growth 39,
45–60.
Chunduri S. K. (2010), ‘New choices for ‘selective’ shoppers’, Photon International,
December, pp. 158–172.
Corkish R. (1991), ‘Some candidate materials for lattice-matched liquid phase epitaxial
growth on silicon’, Solar Cells 31, 537–548.
Cousins P. J., Smith D. D., Luan H. C., Manning J., Dennis T. D., Waldhauer A., Wilson
K. E., Harley G. and Mulligan G. P. (2010), ‘Gen III: improved performance at lower
cost’, 35th IEEE PVSC, June, Honolulu.
De Ceuster D., Cousins P., Rose D., Vicente D., Tipones P. and Mulligan W. (2007), ‘Low
cost, high volume production of >22% efficiency solar cells’, 22nd European Photo-
voltaic Solar Energy Conf., Milan, September, pp. 816–819.
De Clercq K., Frisson L., Szlufcik J., Nijs J. and Mertens R. (1997), ‘Design for man-
ufacturability of a silicon solar cell process based on screen printing by means of
manufacturing science techniques’, Conf. Proc., 14th European Photovoltaic Solar
Energy Conf., Barcelona, pp. 135–138.
De Moor H. H. C., Hoornstra J., Weeber A. W., Burgers A. R. and Sinke W. C. (1997), ‘Print-
ing high and fine metal lines using stencils’, Conf. Proc. 14th European Photovoltaic
Solar Energy Conf., Barcelona, pp. 404–407.
Dietl J., Helmreich D. and Sirtl E. (1981), “‘Solar” silicon’, in Crystals, Growth, Properties
and Applications, Vol. 5: Silicon, Springer-Verlag, pp. 43–107.
Einhaus R., Vazsonyi E., Szlufcik J., Nijs J. and Mertens R. (1997), ‘Isotropic texturing of
multicrystalline silicon wafers with acidic texturing solutions’, Conf. Proc., 26th IEEE
Photovoltaic Specialists Conf., pp. 167–170.
Fang P. H., Ephrath L. and Nowak W. B. (1974), ‘Polycrystalline silicon films on aluminium
sheets for solar cell application’, Appl. Phys. Lett. 25, 583.
Ferrazza F. (1995), ‘Growth and post-growth processes of multicrystalline silicon for pho-
tovoltaic use’, in Pizzini S., Strunk H. P. and Werner J. H. (eds), Polycrystalline Semi-
conductors IV — Physics, Chemistry and Technology, Solid State Phenomena” Trans.
Tech. Publ., Zug, Switzerland.
Fischer D., Keppner H., Kroll U., Torres P., Meier J., Platz R., Dubail S., Selvan J. A.
A., Vaucher N. P., Ziegler Y., Tscharner R., Hof C., Beck N., Goetz M., Pernet P.,
Goerlitzer M., Wyrsch N., Vuille J., Cuperus J. and Shah A. (1997), ‘Recent progress
of the ‘Micromorph’ tandem solar cells’, Conf. Proc., 14th European Photovoltaic
Solar Energy Conf., Barcelona, pp. 2347–2350.
Fischer H. and Pschunder W. (1976), ‘Low cost solar cells based on large area unconven-
tional silicon’, Conf. Record, 12th IEEE Photovoltaic Specialists Conf., Baton Rouge,
November, pp. 86–82.
Godfrey R. B. and Green M. A. (1980), ‘High efficiency silicon miniMIS solar cells —
design and experimental results’, IEEE Trans. Elec. Dev. ED-27, 737–745.
Godlewski M. P., Baraona C. R. and Brandhorst H. W. (1973), ‘Low-high junction theory
applied to solar cells’, 10th IEEE Photovoltaic Specialists Conf., Palo Alto, pp. 40–49.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch03 page 133
Grassman T. J., Brenner M. R., Rajagopalan S., Unocic R., Dehoff R., Mills M., Fraser
H. and Ringel S. A. (2009), ‘Control and elimination of nucleation-related defects in
GaP/Si(001) heteroepitaxy’, App. Phys. Lett. 94, 232106–232106-3.
Green M. A. (1984), ‘Limits on the open circuit voltage and efficiency of silicon solar cells
imposed by intrinsic Auger processes’, IEEE Trans. Electron Devices ED-31, 671–678.
Green M. A. (1987), High Efficiency Silicon Solar Cells, Trans Tech Publications,
Aedermannsdorf.
Green M. A. (1995), Silicon Solar Cells: Advanced Principles and Practice, Bridge Printery,
Sydney.
Green M. A. (1999), ‘Limiting efficiency of bulk and thin film silicon solar cells in the
presence of surface recombination’, Progr. Photovoltaics 7, 327–330.
Green M. A. (2003), ‘General temperature dependence of solar cell performance and impli-
cations for device modelling’, Progr. Photovoltaics 11, 333–340.
Green M. A. (2009), ‘The path to 25% silicon solar cell efficiency: history of silicon cell
evolution’, Progr. Photovoltaics 17, 183–189.
Green M. A. (2011), ‘Ag requirements for silicon wafer-based solar cells’, Progr. Photo-
voltaics 19, 911–916.
Green M. A., Basore P. A., Chang N., Clugston D., Egan R., Evans R., Ho J., Hogg
D., Jarnason S., Keevers M., Lasswell P., O’Sullivan J., Schubert U., Turner A.,
Wenham S. R. and Young T. (2004), ‘Crystalline silicon on glass (CSG) thin-film
solar cell modules’, Solar Energy, Special Issue on Thin Film Photovoltaics 77,
857–863.
Green M. A. and Hansen J. (1998), Catalogue of Photovoltaic Drawings, Photovoltaics
Special Research Centre, University of New South Wales, Sydney.
Green M. A. and Ho J. (2011), Catalogue of Photovoltaic Drawings, 3rd edition, School
of Photovoltaic and Renewable Energy Engineering, University of New South Wales,
Sydney.
Green M. A., Zhao J., Wang A. and Wenham S. R. (1992), ‘45% efficient silicon photovoltaic
cell under monochromatic light’, IEEE Electron Device Lett. 31, 317–318.
Haase C. and Podewils C. (2011), ‘More of everything: market survey on solar modules
2011’, Photon International, February, pp. 174–221.
Hahn G. and Schönecker A. (2004), ‘New crystalline silicon ribbon materials for photo-
voltaics’, J. Phys.: Condens. Matter 16, R1615–R1648.
Hall R. B., Barnett A. M., Brown J. E., Checchi J. C., Ford D. H., Kendall C. L., Mulligan
W. P., Rand J. A. and Ruffins T. R. (1994), ‘Columnar-grained polycrystalline solar
cell and process of manufacture’, U.S. Patent No. 5336335, August 9, 1994.
Haynos J., Allison J., Arndt R. and Meulenberg A. (1974), ‘The COMSAT non-reflective
silicon solar cell: a second generation improved cell’, Int. Conf. on Photovoltaic Power
Generation, Hamburg, p. 487.
Healy S. A. and Green M. A. (1992), ‘Efficiency enhancements in c-Si solar cells by the
incorporation of a region alloyed with germanium’, Solar Energy Mater. Solar Cells
28, 273–284.
Hering G. (2011), ‘Year of the Tiger: PV cell output roared in 2010 to over 27 GW —
beating 2006 through 2009 combined — but can the Year of the Rabbit bring more
multiples?’, Photon International, March, pp. 186–218.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch03 page 134
134 M. A. Green
Janoch R., Wallace R. and Hanoka J. I. (1997), ‘Commercialization of silicon sheet via
the string ribbon crystal growth technique’, Conf. Record, 26th IEEE Photovoltaic
Specialists Conf., Anaheim, September/October, pp. 95–98.
Jordan D. and Nagle J. P. (1994), ‘New generation of high-efficiency solar cells: develop-
ment, processing and marketing’, Progr. Photovoltaics 2, 171–176.
Keevers M. J. and Green M. A. (1994), ‘Efficiency improvements of silicon solar cells by
the impurity photovoltaic effect’, J. Appl. Phys. 75, 4022–4033.
Keevers M. J., Young T. L., Schubert U. and Green M. A. (2007), ‘10% efficient CSG min-
imodules’, 22nd European Photovoltaic Solar Energy Conf., Milan, 3–7 September,
pp. 1783–1790.
Keevers M. J., Zhang G. C., Saris F. W., Zhao J. and Green M. A. (1995), ‘Screening of
optical dopants in silicon solar cells for improved infrared response’, Conf. Proc., 13th
European Photovoltaic Solar Energy Conf., Nice, pp. 1215–1218.
Khattak C. P., Basaran M., Schmid F., D’Aiello R. V., Robinson P. H. and Firester A. H.
(1981), ‘Metallurgical-silicon substrates produced by HEM for epitaxial thin film solar
cells’, Conf. Record, 15th IEEE Photovoltaic Specialists Conf., Orlando, May 12–15,
pp. 1432–1437.
King D. L. and Buck E. M. (1991), ‘Experimental optimization of an anisotropic etch-
ing process for random texturization of silicon solar cells’, Conf. Record, 22nd IEEE
Photovoltaic Specialists Conf., Las Vegas, October, pp. 303–308.
Kingsbury E. F. and Ohl R. S. (1952), ‘Photoelectric properties of ionically bombarded
silicon’, Bell Syst. Tech. J. 31, 8092.
Knobloch J., Glunz S. W., Biro D., Warta W., Schaffer E. and Wettling W. (1996), ‘Solar
cells with efficiencies above 21% processed from Czochralski grown silicon’, Conf.
Record, 25th IEEE Photovoltaic Specialists Conf., Washington, Washington D.C., May,
pp. 405–408.
Kolodinski S., Werner J. H., Wittchen T. and Queisser H. J. (1993), ‘Quantum efficiencies
exceeding unity due to impact ionization in silicon solar cells’, Appl. Phys. Lett. 63,
2405–2407.
Lamers M. W. P. E., Tjengdrawira C., Koppes M., Bennett I. J., Bende E. E., Visser T. P.,
Kossen E., Brockholz B., Mewe A. A., Romijn I. G., Sauar E., Carnel L., Julsrud S.,
Naas T., de Jong P. C. and Weeber A. W. (2011), ‘17.9% metal-wrap-through mc-Si
cells resulting in module efficiency of 17.0%’, Progr. Photovoltaics, published online:
23 Mar 2011, DOI: 10.1002/pip.1110.
Lange H. and Schwirtlich I. A. (1990), ‘Ribbon growth on substrate (RGS) — a new approach
to high speed growth of silicon ribbons for photovoltaics’, J. Cryst. Growth 104,
p. 108.
Laudisio G., Young R., Li Z. and Getty R. (2010), ‘A view of the design challenges involved
in the development of advanced n-type contacts using lead-free chemistries’, Proc. 2nd
Workshop on Metallization for Crystalline Silicon Solar Cells — Status, Trends and
New Directions, 14–15 April, Constance, Germany.
Lindmayer J. (1976), ‘Semi-crystalline silicon solar cells’, Conf. Record, 12th IEEE Pho-
tovoltaic Specialists Conf., Baton Rouge, November, pp. 82–85.
Lindmayer J. and Allison J. (1973), ‘The violet cell: an improved silicon solar cell’, COMSAT
Tech. Rev. 3, 1–22.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch03 page 135
Maruyama E., Terakawa A., Taguchi M., Yoshimine Y., Ide D., Baba T., Shima M., Sakata H.
and Tanaka M. (2006), ‘Sanyo’s challenges to the development of high-efficiency HIT
solar cells and the expansion of HIT business’, 4th World Conf. on Photovoltaic Energy
Conversion (WCEP-4), May, Hawaii.
Minagawa S., Saitoh T., Warbisako T., Nakamura N., Itoh H. and Tokuyama T. (1976),
‘Fabrication and characterization of solar cells using dendritic silicon thin films grown
on alumina ceramic’, Conf. Record, 12th IEEE Photovoltaic Specialists Conf., Baton
Rouge, November, pp. 77–81.
Neuhaus D.-H. and Münzer A. (2007), ‘Industrial silicon wafer solar cells’, Adv. OptoElec-
tronics 2007, Article ID 24521, 1–15.
Ohl R. S. (1941), ‘Light sensitive electric device’, U.S. Patent No. 2,402,622, (27
March), ‘Light-sensitive electric device including silicon’, U.S. Patent No. 2,443,542
(27 May).
Photovoltech (2010), Doc. D-S&M-008, Rev. A, 22 Oct, www.photovoltech.com.
Ralph E. L. (1975), ‘Recent advancements in low cost solar cell processing’, Conf. Proc.,
11th IEEE Photovoltaic Specialists Conf., Scottsdale, AZ, pp. 315–316.
Riordan M. and Hoddeson L. (1997), Crystal Fire: The Birth of the Information Age, Norton,
New York.
Rodriguez H., Guerrero I., Koch W., Endrös A. L., Franke D., Häßler C., Kalejs J. P. and
Möller H. J. (2011), ‘Bulk crystal growth and wafering for PV’, in Luque A. and
Hegedus S. (eds) Handbook of Photovoltaic Science and Engineering, 2nd Edition,
Wiley, New York.
Sawada T., Terada N., Tsuge S., Baba T., Takahama T., Wakisaka K., Tsuda S. and Nakano S.
(1994), ‘High-efficiency a-Si/c-Si heterojunction solar cell’, Conf. Record, 1st World
Conf. Photovoltaic Energy Conversion, Hawaii, pp. 1219–1225
Scaff J. H., Theuerer H. C. and Schumachor E. E. (1949), ‘P-type and n-type silicon and
the formation of the photovoltaic barrier in silicon ingots’, Trans. Amer. Inst. Mining
Met. Engineering, June, pp. 383–388.
Schmidt F. (2011), ‘History of technologies development for silicon cost reduction’, in Palz
W., Power for the World, Pan Stanford, Singapore.
Schmidt J. and Bothe K. (2004), ‘Structure and transformation of the metastable boron- and
oxygen-related defect center in crystalline silicon’, Phys. Rev. B 69, 241071–241078.
Schott (2009), ‘Realignment of wafer business at WACKER SCHOTT Solar’, Press Release,
30 September.
Schulz M. and Sirtl E. (1984), ‘Silicon sheet’ in Landolt-Börnstein: Numerical Data and
Functional Relationships in Science and Technology, Springer-Verlag, Berlin, 17c,
Section 6.1.2.5.3, 52–54 and 442–444.
Schumann M., Orellana Perez T. and Riepe S. (2009), ‘The solar cell wafering process’,
Photovoltaics International, August, 53–59.
SEMI (2011), Crystalline silicon PV technology and manufacturing group, International
Technology Roadmap for Photovoltaics: Results 2010, 2nd Edition, SEMI PV Group
Europe, Berlin, March.
Serreze H. B. (1978), ‘Optimizing solar cell performance by simultaneous consideration of
grid pattern design and interconnect configurations’, Conf. Record, 13th IEEE Photo-
voltaic Specialists Conf., Washington, D.C., pp. 609–614.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch03 page 136
136 M. A. Green
Shah A., Meier J., Torres P., Kroll U., Fischer D., Beck N., Wyrsch N. and Keppner H.
(1997), ‘Recent progress on microcrystalline solar cells’, Conf. Record, 26th IEEE
Photovoltaic Specialists Conf., Anaheim, September/October, pp. 569–574.
Solland (2009), Preliminary Data Sheet, Sunweb, 14 May, www.sollandsolar.com.
Song J., Boas R., Bolman C., Meyers M., Rogol M. and Trangucci R. (2010), ‘The true cost
of solar power: how low can you go?’, Photon Consulting, April.
Stoddard N. G., Wu B., Witting I., Wagener, M. Park Y. Rozgonyi G. and Clark R. (2008),
‘Casting single crystal silicon: novel defect profiles from BP Solar’s Mono2TM wafers’,
Solid State Phenomena 131–133, 1–8.
Strahm B., Andrault Y., Baetzner D., Guérin C., Holmes N., Kobas M., Lachenal D., Mendes
B., Tesfai M., Wahli G., Wuensch F., Buechel A., Mai J., Schulze T. and Vogt M. (2010),
‘Progress in silicon heterojunction solar cell development and scaling for large scale
mass production use’, 25th European Photovoltaic Solar Energy Conf., September,
Valencia, Spain.
Tanaka M., Taguchi M., Takahama T., Sawada T., Kuroda S., Matsuyama T., Tsuda S.,
Takeoka A., Nakano S., Hanfusa H. and Kuwano Y. (1993), ‘Development of a new
heterojunction structure (ACJ-HIT) and its application to polycrystalline silicon solar
cells’, Progr. Photovoltaics 1, 85–92.
Tiedje T., Yablonovitch E., Cody G. D. and Brooks B. G. (1984), ‘Limiting efficiency of
silicon solar cells’, IEEE Trans. Electron Devices ED-31, 711–716.
Trupke T., Green M. A. and Würfel P. (2002a), ‘Improving solar cell efficiencies by the
up-conversion of sub-band-gap light’, J. Appl. Phys. 92, 4117–4122.
Trupke T., Green M. A. and Würfel P. (2002b), ‘Improving solar cell efficiencies by down-
conversion of high-energy photons’, J. Appl. Phys. 92, 1668–1674.
Tsao C.-Y. (2011), ‘Fabrication and characterization of Ge thin films and Ge-rich SiGe alloys
for photovoltaic applications’, PhD Thesis, University of New South Wales, submitted
on 31 March.
Wallace R. L., Hanoka J. I., Narasimha S., Kamra S. and Rohatgi A. (1997), ‘Thin silicon
string ribbon for high efficiency polycrystalline solar cells’, Conf. Record, 26th IEEE
Photovoltaic Specialists Conf., Anaheim, September/October, pp. 99–102.
Wenham S., Robinson R., Dai X., Zhao J., Wang A., Tang, Y. H., Ebong A., Honsberg
C. B. and Green M. A. (1994), ‘Rear surface effects in high efficiency silicon solar
cells’, Conf. Record, 1st World Conf. on Photovoltaic Energy Conversion, December,
pp. 1278–1282.
Wenham S. R. and Green M. A. (2002), ‘Self aligning method for forming a selective emitter
and metallization in a solar cell’, U.S. Patent No. US6,429,037 B1, 6 August.
Werner J. H., Brendel R. and Queisser H. J. (1994), ‘New upper efficiency limits for semi-
conductor solar cells’, Conf. Record, 1st World Conf. Photovoltaic Energy Conversion,
Waikoloa, December, pp. 1742–1745.
Wistey M. A., Fang Y.-Y., Tolle J., Chizmeshya A. V. G. and Kouvetakis J. (2007), ‘Chemical
routes to Ge/Si(100) structures for low temperature Si-based semiconductor applica-
tions’, Appl. Phys. Lett. 90, 82108.
Wolf M. (1960), ‘Limitations and possibilities for improvement of photovoltaic solar energy
converters’, Proc. IRE 48, 1246–1263.
Wolf M. (1976), ‘Historical development of solar cells’, in Solar Cells, Backus C. E., ed.,
New York, IEEE Press.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch03 page 137
Yamamoto K., Yoshimi M., Suzuki T., Okamoto Y., Tawada Y. and Nakajima A. (1997),
‘Thin film poly-Si solar cell with ‘Star Structure’ on glass substrate fabricated at
low temperature’, Conf. Record, 26th IEEE Photovoltaic Specialists Conf., Anaheim,
September/October, pp. 575–580.
Zhao J., Wang A., Altermatt P. and Green M. A. (1995), ‘24% efficient silicon solar cells
with double layer antireflection coatings and reduced resistance loss’, Appl. Phys. Lett.
66, 3636–3638.
May 2, 2013 14:6 BC: 8831 - Probability and Statistical Theory PST˙ws
CHAPTER 4
One advantage of being disorderly is that one is constantly making exciting discoveries.
A. A. Milne.
4.1 Introduction
Since the first deposition of hydrogenated amorphous silicon out of a silane dis-
charge (Chittik et al., 1969), plasma coating of thin silicon layers has played
an increasingly important role for a large number of applications in ‘macro-
electronics’, including flat panel displays (FPD), solar modules and X-ray imagers.
Indeed, thin silicon layers acting as active semiconductors were likely coated over
more than 140 million square metres in 2010, on glass plates, metal foils and plas-
tic webs. Hydrogenated amorphous silicon used as an active semiconductor n-type
channel in thin-film transistors (TFT) in active-matrix liquid crystal displays made
the bulk of this surface, whereas around 6% of the surface was dedicated to photo-
voltaic (PV) applications. The recent synergy between the FPD sector and thin-film
silicon PV allowed module producers for the first time in thin-film PV history to
purchase thin-film turnkey lines for large area (> 1 m2 ) coating. Manufacturing
costs below 0.5 /Wp (Kratzla et al., 2010; Ringbeck and Stutterlueti, 2012) with
10% module efficiency or higher are now demonstrated by several companies that
provide equipment, making thin-film silicon a low-cost technology, even at rela-
tively low production volume. Short-term perspective cost in the range of 0.35 /Wp
139
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch04 page 140
are even indicated. In a growing photovoltaic market, which could reach several
hundreds of GW annually, thin-film silicon should be able to remain a player
because of many inherent advantages that are the subject of this chapter.
The key feature of ‘thin-film silicon’ layers for photovoltaic application is
the possibility to manufacture a wide variety of materials with similar processes,
typically by plasma-enhanced chemical vapour deposition (PECVD):
Table 4.1 summarises the various forms of silicon films as they are used today in
PV and FPD on industrial levels. This chapter will focus on the class of amorphous
and microcrystalline materials used as absorber and doped layers in thin-film solar
cells, as well as the related technological aspects. Notably, such layers can be
made with a variety of electronic properties and nanostructures, which depend
on the deposition process and on the substrate (i.e. its surface chemistry and its
geometry).
The chapter is organised as follows. The first section gives an overview of
device configuration, technology and history. The layers’ basic properties are
detailed in Section 4.2. Section 4.3 presents the elementary properties of amorphous
and multijunction thin-film silicon solar cells. Section 4.4 focuses on device-grade
absorber deposition of a-Si:H and µc-Si:H materials. In Section 4.5, aspects related
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch04 page 141
Table 4.1 Silicon-based materials used in thin-film Si solar cells and transistors.
Material Application Typical properties In production?
to light trapping and rough front (transparent conducting oxide) and back electrodes
(e.g. metallic back reflectors) are presented. Section 4.6 reviews some of the newest
device architecture based on novel transparent conducting oxide and doped layers.
Finally, in Section 4.7 up-scaling, production costs and environmental impacts of
the technology are discussed.
ZnO buffer
back ZnO Metal contact
dielectric
reflector
All schemes in Figure 4.1 show rough interfaces within the devices. These
rough interfaces are essential to ensure efficient light scattering inside the silicon
layers and thus to increase the effective light path. The light management schemes
that are used in thin-film silicon technology are discussed in Section 4.5, while
advanced concepts and substrates are presented in Section 4.6. The properties of the
front TCO and back reflector are essential for achieving high device performance;
their optimisation is as important as the PECVD deposition process of the silicon
films. Two important aspects emerge for the choice of substrate texture:
• In superstrate devices, textured TCOs control the maximum current density
via their transmission and light scattering potential.
• The texture of the growth template can lead to defects in the devices, such as
collision of growth fronts, which can result in reduced electrical performance.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch04 page 144
large solar parks in high irradiance areas or for integration of aesthetically pleasing
and safe PV products into buildings, but it certainly needs further increased effi-
ciency in the medium term to stay competitive with other mainstream technologies.
• Figure 4.3 shows that the bandgap of amorphous silicon is between 1.7 to
1.8 eV; this is generally considered to be too high for efficient absorption of
the solar spectrum. Attempts were therefore made to lower the bandgap by
alloying with germanium or tin, or to find new materials with lower gaps for
the construction of advanced tandem and multijunction devices.
• While the demonstration of doping was surprising because an amorphous
network should easily incorporate elements with lower or higher valence, it was
soon noted that the achievable doping efficiency is limited; films with better
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch04 page 147
6
10
Absorption coefficient (cm-1)
5
10
4
10
3
10
2 a-Si:H
10
c-Si:H
10
1 c-Si
0
10
-1
10
1.0 1.5 2.0 2.5 3.0 3.5 4.0
Photon energy (eV)
Figure 4.3 Spectral absorption of the thin-film materials a-Si:H and µc-Si:H (open squares
and circles, respectively), and c-Si (filled circles). Dashed lines illustrate the Urbach tail
between 1.25 and 1.75 eV and the indirect bandgap for energies beyond 1.75 eV.
conductivity became eventually available, but they were found to contain small
crystallites of silicon.
• Such crystallite-containing silicon is actually a material in its own right, called
microcrystalline or nanocrystalline silicon, useful not only for doped layers
but also as an absorber with a low bandgap. Figure 4.3 shows that its absorption
is similar to that of crystalline silicon, i.e. it extends into the near IR region,
but it is also relatively weak (101 –103 cm−1 between 1.5 and 1.12 eV).
• Absorption in the doped layers of solar cells does not contribute to the photo-
current. Recombination losses in doped amorphous layers are unavoidable
because defects are inherently created along with doping. In solar cell
applications, losses in the doped layers were eventually reduced by using
materials with higher bandgap, e.g. by incorporation of carbon, oxygen or
nitrogen.
In the remainder of this section these and other aspects will be addressed in more
detail.
23
10
22
Density of states (1/cm3eV) 10
21
10 VB-tail CB-tail
20
10
19
10
18 D+ D0 D-
10
17
10
16
10
-0.5 0.0 0.5 1.0 1.5 2.0
Energy with respect to VB edge (eV)
Figure 4.4 Schematic density of states in device-grade a-Si:H. For recombination statistics,
often two transition states D+ /D0 and D0 /D− are used rather than actual charge states.
defects must be taken into account by the device design, e.g. by charge transport
via drift rather than diffusion.
4.2.3 Doping
Doping of a-Si:H with phosphine (PH3 ) to make n-type material was already
reported with the first demonstration of a-Si:H (Chittick et al., 1969). Later, p-type
doping was also achieved by adding diborane (B2H6 ) to a glow discharge (Spear and
Le Comber, 1976), but the doping efficiency in amorphous silicon always remained
low. The assumption of amphoteric defects in the middle of the bandgap can explain
this behaviour: if n-type doping moves the Fermi level towards the conduction band,
the defect eventually lies below the Fermi level. It can be charged with a second
electron after overcoming the correlation energy U . Likewise, if p-type doping
moves the Fermi level towards the valence band, the unpaired electron residing
in the neutral defect D0 is above the Fermi level and it becomes energetically
more favourable to form a charged defect D+ . Free carriers of either type are thus
absorbed by the defect state, putting a stop to any further doping.
Figure 4.5 illustrates the energy gain by the two processes, assuming a bandgap
of 1.75 eV, a defect state at E D = 0.85 eV above the valence band, and a correlation
energy of U = 0.2 eV. The predicted defect densities according to the equilibrium
model and some experimental data are also illustrated in Figure 4.5; whenever
the Fermi level is moved too far from mid-gap, the formation energy of one type
Formation energy (eV)
1.0
0.8
D0
0.6
0.4 D+ D-
0.2
0.0
18
10
17
10
ND (cm-3)
16
10 p-type n-type
15
10
14
10
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6
Fermi level w.r. to VB (eV)
Figure 4.5 Formation energies of defects in their various charge states with respect to
the Fermi level (upper panel). The lower panel compares the predicted equilibrium defect
densities, squares and circles refer to reported data of p- and n-doped samples, respectively.
Data after Winer (1990).
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch04 page 152
of charged defect starts dropping below the level of that of the neutral defect.
Consequently, the formation of this defect becomes more likely and works against
any further doping. The lowest possible defect concentrations are found in a narrow
region where the neutral defect dominates.
Note that shifts of the Fermi level are not restricted to doped materials, but also
occur throughout the intrinsic layer of a p-i -n junction. The equilibrium models
thus predict increased densities of positively and negatively charged defects close
to the p-i and i -n interfaces, respectively. Charging of the defects close to the p-i
interface gives rise to the band bending observed in Figure 4.7. The understanding
of these effects led to the introduction of buffer layers and bandgap grading close
to the interfaces where the defect densities are projected to be the highest (Guha
et al., 1989).
While the equilibrium idea is quite powerful, there is still controversy whether
such a comparatively simple model with a single defect state and three different
charge states is sufficient. Indeed, to date no model can explain consistently how
bond breaking and formation takes place on a microscopic level. Also, evidence
from cell degradation suggested early on that there is a difference between fast and
slow defects (Yang and Chen, 1993). In addition, defect densities determined by
two independent methods, such as lifetime and defect absorption measurements,
have repeatedly shown very different kinetics during degradation-annealing cycles
(Stradins, 2003), suggesting the existence of various defect states rather than iso-
lated dangling bonds.
more stable monohydrides starts only beyond 450◦ C. In the previous section it was
shown that bond breaking is related to the position of the Fermi level. This explains
the observation that hydrogen effusion depends on doping (Beyer, 2003). Silicon-
hydrogen bonds also have distinct signatures in Fourier transform infrared (FTIR)
spectroscopy; stretching modes that absorb at 2010 cm−1 are generally attributed to
monohydride bonding of dense material whereas a shift of this signature towards
2090 cm−1 is attributed to di- or tri-hydrides which are often found in porous
material.
While the equilibrium models discussed in Sections 4.2.2 and 4.2.3 were
developed around the idea of a continuous random silicon network with isolated
defects, experimental evidence suggests that hydrogen can be bonded also at inter-
nal surfaces of nano-sized voids that exist even in device quality material (Mahan
et al., 1989). This is corroborated by measurements of mass density and hydrogen
content that are compatible with the mass deficit of hydrogenated di-vacancies
over a wide range of deposition conditions (Smets et al., 2007a). In this context,
the above-mentioned IR absorption signatures are attributed to shifted resonance
frequencies of hydrogen atoms that are confined to anisotropic volume defects such
as hydrogenated di- and poly-vacancies (1980–2010 cm−1, dense material), or to
quasi-free oscillations of hydrogen located on the surface of microscopic voids
(2090 cm−1, porous material). The integrated intensities of the two contributions
are used to define the microstructure factor R = I2090 /(I2090 + I2010 ) for distin-
guishing dense and porous material, but more importantly, a dominant presence of
monohydride bonding is also associated with reduced defect density (Kroll et al.,
1998).
Figure 4.6 Schematic transition from fully amorphous silicon (left) over mixed-phase mate-
rial to fully microcrystalline silicon (right) with increasing hydrogen content in the precursor
gas mix.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch04 page 155
2
TCO <p> <i> <n>
1
Potential (eV)
CB
VB
-1
<p> <i>
-2
0 50 100 150 200
Distance from TCO/p interface (nm)
Figure 4.7 Equilibrium band diagram of an amorphous silicon solar cell with a 200 nm thick
absorber layer and p- and n-layers with activation energies of 450 and 250 mV, respectively.
Potentials are shown with respect to the Fermi level. Dashed and solid lines represent ideal
and realistic band alignment, respectively. The inset illustrates the Schottky junction between
the front TCO and the p-layer.
1.0
0.8
0.6
EQE
0.4
2
14.5 mA/cm
2
0.2 13.6 mA/cm
0.0
400 500 600 700 800
Wavelength (nm)
Figure 4.8 EQE of an n-i- p a-Si:H single-junction cell without antireflection coating. Solid
and dashed lines illustrate the charge-carrier collection in the initial and stabilised states,
respectively.
by the illumination spectrum (in units of photons per wavelength interval and
per surface area), and integrating over all wavelengths. This determination of the
short-circuit current density from the EQE is normally more precise than the value
obtained from a current–voltage measurement because the EQE does not depend on
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch04 page 158
the cell area, whereas a reliable determination of the cell area can become difficult
for small cells, particularly if they include light-scattering interfaces.
After LID, the drift field is weakened and the collection of photogenerated
charge carriers becomes even less effective under the forward bias of operating
conditions. As a result, the solar cell efficiency is further reduced via losses in fill
factor (FF). Intuitively, Voc should be related to the potential created by the doped
layers and thus be minimally influenced by LID. However, this simplistic view
does not take into account that Voc can suffer because of increased recombination
once the layers become more defective.
Figure 4.9 illustrates the kinetics of cell efficiency degradation. The degrada-
tion scales with the absorber layer thickness: thinner layers degrade less because
the remaining field in the centre of the junction remains stronger in thinner cells,
even when weakened by additional defects. Moreover, the figure shows that cells
in which the i -layer is deposited with additional hydrogen stabilise faster and at
a higher efficiency for a given thickness (Wronski, 1996). In general, there is no
simple dependence on illumination duration, intensity or spectrum, and even at
moderate temperatures of 60 or 70◦C, the degradation is partly reversed by anneal-
ing. Therefore, standard degradation experiments are commonly carried out for
1000 hours at 50◦ C under exposure to light with the intensity and the spectrum of
sunlight. Reduced cell thickness is desirable, not only in terms of stability against
LID, but also because of throughput considerations in industrial production. A
sufficient level of light absorption must nevertheless be maintained. Absorption
1.0
85 nm
0.9
Normalised efficiency
0.8
0.7 240
nm
0.6 40
0n
m
0.5 64
0n
m
0.4
1 10 100 1000
Degradation time (h)
Figure 4.9 Degradation kinetics of solar cells with various absorber layer thicknesses; data
from Bennett et al. (1986) and Bennett and Rajan (1988). The solid line represents a cell
with a 400 nm i-layer deposited with hydrogen dilution. Data from Wronski (1996).
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch04 page 159
• Historically, the first tandem cells combined two purely amorphous silicon
absorbers; while they do not extend the spectral response compared with a
single-junction cell, they are nevertheless advantageous because of reduced
degradation with respect to a single-junction cell of comparable thickness
(Hanak and Korsun, 1982; Bennett and Rajan, 1988; Lechner et al., 2008).
Figure 4.11 shows the repartition of 15 mA cm−2 between a thin a-Si:H top
and a relatively thick a-Si:H bottom cell.
• The combination of amorphous and microcrystalline cells in a tandem with
high total current appears unfavourable because the top cell sees only a single
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch04 page 160
µc-Si:H bo om cell
SiOx int. reflector
a-Si:H top cell
1 µm µc-Si:H bo om cell
Figure 4.10 Cross-sections through micromorph tandem cells with intermediate reflectors.
The upper panel shows a p-i-n/p-i-n structure on glass covered with a 4 µm thick ZnO front
contact. The intermediate reflector layer appears as a light band between the thin amorphous
and the thick microcrystalline films. The lower panel shows an n-i-p/n-i-p tandem cell on a
periodically textured plastic substrate. In this cell, the intermediate reflector is about 2 µm
thick. The growth sequence of both cells is from bottom to top.
pass of light and the resulting tandem would thus be heavily mismatched. This
limitation was successfully overcome by the introduction of a ZnO intermedi-
ate reflector layer between the component cells, resulting in much better top
cell current (Fischer et al., 1996; Yamamoto et al., 2002; Yamamoto et al.,
2006). An alternative material for the intermediate reflector is silicon oxide
doped with phosphorous (SiOx :P) (Buehlmann et al., 2007). An advantage
of SiOx :P is the possibility of depositing it between silicon layers without
breaking vacuum; this is discussed in more detail in Section 4.6.1.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch04 page 161
1.0
1.0
13.8 mA/cm2
0.8 0.8
13.9 mA/cm2
0.6 0.6
EQE
EQE
0.4 top: bottom:
0.4
70 nm 300 nm
2 2
7.30 mA/cm 7.35 mA/cm
0.2 0.2
0.0 0.0
400 500 600 700 800 400 500 600 700 800 900 1000 1100
Wavelength [nm] wavelength (nm)
Figure 4.11 EQEs of an a-Si:H/a-Si:H tandem cell (left) and of a micromorph cell with an
intermediate reflector layer (right). The micromorph cell has a 240 nm a-Si:H top cell, an
80 nm SiOx intermediate reflector, and a 2.4 µm thick bottom cell. The currents represent
the integrated i sc of the sub-cells. More information on these results can be found in Cubero
et al. (2008) and Boccard et al. (2011), respectively.
Figure 4.12 EQEs of triple-junction cells. The left panel represents a fully amor-
phous a-Si/a-SiGe/a-SiGe cell (Yang et al., 1997), the right panel represents an a-Si:H/
µc-Si:H/µc-Si:H cell with AR-coating (Söderström, 2013). The currents represent the inte-
grated current densities of the sub-cells.
• Thanks to the continuous variation of bandgaps in the a-Six Ge1−x :H alloy sys-
tem, a 10 mA cm−2 a-Si top cell can easily be combined with an a-Si x Ge1−x :H
bottom cell. Cell design must maximise stabilised efficiency by trading off the
germanium content vs. bottom cell thickness trade-off. Going to triple-junction
cells, somewhat higher germanium contents are admissible because of the
further reduced thicknesses of the component cells. Figure 4.12 shows that
triple-junction a-Si:H/a-Six Ge1−x /a-Si y Ge1−y cells can be matched at about
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch04 page 162
8.3 mA cm−2 (Guha et al., 2000). In the last two approaches no intermediate
reflector is required below the a-Si:H top cell.
• An alternative approach distributes the high current of a microcrystalline sil-
icon cell into a triple-junction cell consisting of an amorphous top cell and a
microcrystalline tandem bottom (a-Si:H/µc-Si:H/µc-Si:H). Using a slightly
thinner and therefore more stable top cell yields current matching between 9
and 10 mA cm−2 (Saito et al., 2005; Yue et al., 2010; Söderström, 2013). The
potential of going to efficiencies beyond 16% with this approach is discussed
in Section 4.6.6.
n-type doping, respectively. A voltage is then applied to the parallel electrodes; free
electrons present in the reactor are accelerated and gas molecules are dissociated
by electron impact ionisation, generating more free electrons which trigger further
impact ionisations until the discharge reaches a stable state. Radio frequency (RF)
excitation predominantly increases the energy of the light electrons while the heavy
ions cannot follow the rapid switching of the alternating field. Thus, RF-PECVD
has the advantage of accessing high-temperature chemistry while remaining at
typical substrate temperatures of 150 to 250◦ C (cold plasma).
Electrons have much higher thermal velocities than ions and can thus reach
the electrodes faster. As a consequence, electric fields develop near the electrodes
which retard electrons and accelerate ions. These regions in front of the electrodes
are called sheaths. They extend over small distances (in the mm range at normal
deposition pressures) and drive the ion bombardment of the surfaces. The standard
RF frequency is 13.56 MHz (as allotted for industrial processes by international
authorities). Alternatively, very high frequencies (VHF, 20–150 MHz) proved use-
ful for achieving higher deposition rates (Curtins et al., 1987; Kroll et al., 1997)
and are today widely used in the thin-film silicon PV industry. Finally, microwave
excitation (2.54 GHz) is also being used in electron-cyclotron resonance PECVD
(ECR-PECVD) (Bae et al., 1998) and for inline microwave PECVD remote plasma
sources (Soppe et al., 2005).
In the plasma bulk, electron impact dissociation of SiH4 and H2 molecules
takes place, producing silicon radicals, SiHx (x = 1–3) and atomic hydrogen, H.
Radicals, molecules and ions can further interact in secondary reactions, such as
higher silane formation from the iterative insertion reaction SiH2 + Sin Hm →
Sin+1 Hm+2 , which triggers powder formation (Bano et al., 2005). Radicals diffuse
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch04 page 164
from the plasma zone where they are generated towards the reactor walls, the elec-
trodes and the substrate. Growth takes place under a complicated non-equilibrium
process that is governed by the temperature-dependent sticking coefficients, surface
diffusivity of silicon radicals, etching processes by atomic hydrogen, and surface
and sub-surface hydrogen recombination. Under adequate conditions, a silicon film
with a low hydrogen content of around 10 at.% is deposited.
model, high H flux on the surface enhances the mobility of the deposition precur-
sors, strongly influencing the crystalline fraction of the deposited film (Matsuda
et al., 1999). In the selective etching model, amorphous and crystalline phases are
deposited simultaneously with H preferentially etching the amorphous material,
leaving behind the crystallites (Terasa et al., 2000). Finally, the chemical annealing
model postulates that H atoms induce crystallisation of amorphous films (Sriraman
et al., 2002). These three models imply that a high flux of atomic hydrogen H
with respect to the flux of silicon radicals Si favours microcrystalline silicon
growth. A well-known way to reach the transition region between amorphous and
microcrystalline silicon is therefore to use silane that is highly diluted in hydrogen,
but a high H /Si ratio can also be attained from pure silane plasma if sufficient
H is provided by high silane dissociation efficiency. The H /Si ratio required to
reach the transition zone was experimentally assessed (Dingemans et al., 2008)
and also related to the plasma silane concentration c p (Strahm et al., 2007); this
latter parameter is defined by the input silane concentration c and by the deple-
tion fraction of silane in the plasma D as c p = c(1−D). The transition zone was
shown to occur at 0.5% < c p < 1.2%; higher (lower) c p systematically results in
amorphous (microcrystalline) layers.
Figure 4.14 illustrates two optimum deposition windows for the production
of high-quality materials at the transition. In the low silane concentration regime
Figure 4.14 Left: Amorphous to microcrystalline transition zone (shaded region) for vary-
ing silane concentration and depletion fraction, with low silane concentration labelled as
Regime 1 and high silane depletion as Regime 2 (Strahm et al., 2007); right: deposition rate
of device-grade microcrystalline material versus the employed power density (data collected
from different institutes) from low pressure (<1 Torr) RF and VHF regimes (triangles and
squares) to high pressure (>1 Torr) RF and VHF regimes (circles and stars) (Smets et al.,
2008).
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch04 page 166
(c ≤ 2.5%), transition is achieved for low depletion (D < 0.5): the film microstruc-
ture can be controlled with the silane-to-hydrogen gas flow ratio and high depo-
sition rates can be achieved only with the use of high gas flows and high power
densities. In the high silane depletion regime (D > 80%), growth in the transition
region can be achieved only for higher silane concentration (c > 5%). This regime
provides a higher efficiency of silane utilisation and lower hydrogen flows, while
the film structure is determined by the depletion which is, for example, accessi-
ble by controlling the plasma power. It must be noted that high-quality materials,
as assessed by FTPS and FTIR measurements, can be produced in the different
process windows presented in this section, and that device-grade materials typi-
cally have a Raman crystalline fraction between 50% and 70%. However, it was
recently demonstrated that similar high-quality bulk materials, as determined by
FTPS and FTIR measurements, can lead to varying device efficiencies (Bugnon
et al., 2011). To realise high-efficiency devices, the plasma process conditions must
also provide a dense material when the deposition is carried out on a textured sub-
strate in order to limit the creation of localised defects (typically 2D nano-porous
material regions) in the sharp valleys of a highly textured substrate. Very impor-
tantly, the most efficient process conditions for high-quality devices must therefore
provide not only high-quality bulk material but also a minimum density of localised
defects or cracks when deposition is carried out on a textured substrate (Bugnon
et al., 2011; Despeisse et al., 2011).
there are no ions and electric fields present and consequently there is no ion bom-
bardment of the growing surface. This is particularly important when depositing
the first part of the intrinsic layer in p-i -n microcrystalline junctions, in order to
avoid damaging the critical p-i interface. The use of intrinsic µc-Si:H buffer lay-
ers deposited by HWCVD at the p-i interface of µc-Si:H junctions resulted in
very high Voc (600 mV) and efficiency (Mai et al., 2005). While this deposition
technique is used on an experimental scale for depositing high-quality materials,
high deposition rates (>10 nm s−1) are also achievable (Mahan et al., 2002). How-
ever, the performance of HWCVD-deposited solar cells has not yet surpassed that
of PECVD-deposited solar cells. Moreover, this deposition technique has not yet
been successfully incorporated into large-scale manufacturing facilities, because
of problems with deposition uniformity and filament aging. To date, HWCVD has
thus been employed only at the laboratory scale for thin-film silicon.
glass
TCO
silicon
back reflector
with respect to the interface is beyond the critical angle of total internal reflection.
Textured interfaces thus multiply the effective absorption length of absorber layers
beyond the maximum of two passes through a flat device.
Considering a slab of weakly absorbing material with completely random
scattering interfaces, Yablonovitch proposed an upper theoretical limit of 4n r2 for
the maximum achievable absorption enhancement, n r being the refractive index
of the absorber material (Yablonovitch and Cody, 1982). It was proposed that
deterministic grating structures can exceed the 4nr2 limit in certain wavelength
ranges (Sheng et al., 1983; Gee, 2002; Yu et al., 2010). However, in solar cells,
the experimentally found path enhancement is normally between 15 and 20, much
below the value of 4n r2 of nearly 50 when a refractive index of 3.5 is assumed for
silicon. A possible explanation for this discrepancy is the neglect of the supporting
layers in most of the theoretical derivations (Haug et al., 2011).
For scattering at real surfaces, a detailed explanation of the light trapping pro-
cess is far from straightforward. The interaction of electromagnetic radiation with
matter can be treated with numerous software packages, but the random nature of
typical textures such as those shown in Figure 4.16 is difficult to describe rigor-
ously. Most approaches revert to super-cells, i.e. box-shaped regions that contain
the multi-layer stack of the solar cell above a base area of several µm2 (Rockstuhl
(a) (b)
(c) (d)
Figure 4.16 Surface morphologies of TCOs with small feature size for a-Si:H solar cells
(top) and larger features for microcrystalline cells (bottom). (a) and (c) show AP-CVD-
grown SnO2 :F of the Asahi-U and Asahi-W type, respectively (Taneda et al., 2007); (b) and
(d) show LP-CVD grown ZnO:B films. The growth processes are described in Faÿ et al.
(2005) and Faÿ et al. (2007). Scale bar: 1 µm.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch04 page 170
et al., 2011). Care must be taken to avoid artefacts because most calculation rou-
tines assume periodic repetition of the cells. The required computing power has so
far limited the use of rigorous models as predictive tools for optimisation; at the
time of writing the calculation of an a-Si:H cell can take hours and a full tandem
cell may take days.
Approximate methods are typically based on scalar scattering theory which
neglects the vectorial nature of electromagnetic fields. Initially developed for
remote sensing of ocean surfaces with radar waves, the scaling to visible light and
nano-textures proved very successful for solar cells (Zeman et al., 2000; Poruba
et al., 2000; Krč et al., 2003). Simulation resembles a ray-tracing approach where
scattering at each interface is described by distribution into a specular and diffuse
part for both reflection and transmission. The diffuse part is normally referred to
as haze; besides its magnitude, a description of the angular intensity distribution
is required (ARS, angle resolved scattering) which is used to describe the average
path enhancement. In scalar scattering theory, the haze depends only on a single
parameter, the root mean square roughness, which is accessible by atomic force
microscopy or optical measurements (Carniglia, 1979; Elson and Bennett, 1979).
Recently, more detailed scattering models based on Fourier theory have been pro-
posed (Dominé et al., 2010).
Adequate surface textures for light management are commonly incorporated
into one of the supporting layers, e.g. the front TCO for p-i -n devices on rigid
glass substrates. Typical examples of fluorinated tin oxide (SnO2 :F) grown by
atmospheric pressure chemical vapour deposition (APCVD) and boron-doped ZnO
(ZnO:B) deposited by low pressure CVD (LPCVD) are shown in the scanning elec-
tron microscope (SEM) images in Figure 4.16. When deposited under conditions
that favour the preferential growth of certain low-energy surfaces, these TCO mate-
rials develop facets that resemble small triangular pyramids. The feature size of
the pyramids can easily be controlled by the film thickness.
Surfaces that are dominated by pronounced V-shaped valleys between the
facets have been found to be detrimental to silicon film growth. Cross-sectional
analysis of amorphous solar cells revealed the growth of defective material above
such valleys because of reduced adatom mobility (Sakai et al., 1990). Figure 4.18
shows that this effect is pronounced in microcrystalline cells, resulting in severe
Voc and FF losses (Nasuno et al., 2001a; Li et al., 2008; Python et al., 2008). The
defective zones have also been identified to provide diffusion paths for oxygen and
other impurities (Python et al., 2010).
Figure 4.17a shows a surface where the initially pyramidal texture of the
LP-CVD ZnO shown in Figure 4.16d has been modified by a plasma treatment that
converts the V-shaped valleys between the facets into rounded valleys with smooth,
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch04 page 171
(a) (b)
Figure 4.17 Surfaces with rounded features. (a) corresponds to the surface shown in Fig-
ure 4.16d after plasma treatment (Bailat et al., 2006); (b) illustrates the crater lime etch pits
of sputtered ZnO etched for 30 s in diluted HCl (Kluth et al., 1999). Scale bar: 1 µm.
Figure 4.18 Top: top-view SEM images of front electrodes (three TCOs with shapes mod-
ified from V to U by plasma treatment); bottom: TEM cross-sections of microcrystalline
silicon layers deposited on the TCOs. There is a clear correlation between the ZnO surface
morphology and the formation of defective zones within microcrystalline absorber films.
Python et al. (2008) give more details.
the crater-like texture of etch pits obtained on sputtered ZnO films after etching in
diluted hydrochloric acid (HCl) (Kluth et al., 1999).
An approach that completely separates the texture for light-confinement from
the growth process has been suggested for the n-i - p structure (Sai et al., 2011).
Starting from a flat silver reflector covered with ZnO, a relief grating is etched
into the ZnO and filled with n-doped amorphous silicon. Subsequently, the stack
is polished in order to provide a flat growth template for the microcrystalline cell.
Owing to the complete absence of valleys that induce the growth of defective zones,
the resulting device showed the same high voltage as a flat reference cell. However,
the n-doped filling-layer resulted in parasitic absorption between 650 and 750 nm.
An alternative design with an undoped filling-layer in combination with a triple-
junction device successfully overcame this limitation (Söderström et al., 2012a;
Söderström, 2013).
Traditionally, n-i - p solar cells relied on the natural textures that develop when
metals like aluminium or silver are deposited on heated substrates (Hirasaka et al.,
1990; Banerjee and Guha, 1991). Because of its relative inertness, the noble metal
silver offers an additional handle on texturing if small quantities of aluminium
are added to the silver sputtering target: by adding oxygen to the gas mixture
during sputtering from such an alloy target, the interface texture is modified by
the inclusion of alumina (Al2 O3 ) crystallites in the growing film (Franken et al.,
2007). The formation of the texture typically requires temperatures of around 300
to 400◦ C; flexible substrate materials are therefore limited to stainless steel (Baner-
jee and Guha, 1991) or polyimides (Takano and Kamoshita, 2004). However, in
order to fully exploit the promise of reduced production cost with roll-to-roll
processing, truly low-cost substrates are desirable. Substrates like polyethylene
thus require a different strategy where the substrate itself is textured; this can
be achieved by hot embossing (Fonrodona et al., 2005) or by imprinting into a
UV-curable polymer (Bailat et al., 2005; Söderström et al., 2010). The interface
texture is then transported into the device by conformal growth. An example of
a microcrystalline n-i - p cell on a periodic substrate texture obtained by roll-to-
roll embossing is shown in the transmission electron microscope (TEM) image
in Figure 4.10; Figure 4.19 shows in more detail how the moderate texture of the
grating substrate is perfectly reproduced by the sputtered silver and ZnO back con-
tact. However, throughout the microcrystalline absorber layer, the initial sinusoidal
texture develops into a structure with a much larger radius of curvature and pinched
valleys.
This change of texture is an issue for tandem cells because in the n-i - p config-
uration the amorphous sub-cell is deposited after the microcrystalline cell. While
the large features of the grating substrate can deliver acceptable isc of about 23 to
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch04 page 173
p-layer
i-layer
n-layer
ZnO /Ag
polymer substrate
25 mA cm−2 in bottom cells with thickness varying between 1.2 and 2 µm, the flat-
tening of the interface texture is clearly inadequate for the top cell. The introduction
of a thick, textured intermediate reflector can ameliorate this situation: thanks to
the texture that ZnO develops when deposited by LP-CVD, Figure 4.10 shows
that the top cell can be grown on an adapted texture that is completely independent
from the bottom cell (Söderström et al., 2009). Tandem cells in n-i - p configuration
with stabilised efficiency up to 11.6% have been demonstrated with this approach
(Biron et al., 2013).
of the first thin-film silicon modules on glass with stabilised efficiencies surpassing
the 10% barrier (Klein et al., 2010; Kratzla et al., 2010). Modules with up to 13.1%
initial efficiency and 10.8% stabilised efficiency were demonstrated in 2011 (Kluth
et al., 2011).
On metal foil, 12% initial efficiency and 11.2% stabilised efficiency were
reported (Banerjee et al., 2011a) for a 400 cm2 triple-junction cell (a-Si:H/a-
SiGe:H/µc-Si:H). These results demonstrate that major bricks for efficient mass
production over large areas are in place. To further increase the efficiency, and thus
the impact, of thin-film silicon technology, advanced light management schemes,
novel nanomaterials and innovative device architectures are currently being devel-
oped. Some of these recent advances are reviewed in this section, together with the
best efficiencies obtained so far at the cell level.
Figure 4.20 Left: SEM image of a focussed ion beam cross-section of a micromorph
tandem cell with LP-CVD ZnO front and back contacts, a 250 nm thick a-Si:H top cell, a
40 nm thick silicon-rich silicon oxide IRL and a 1.1 µm thick bottom cell. The nanometre-
wide filamentary structure of the silicon phase in the SiOx material is illustrated in the
inset EFTEM image. Courtesy of Peter Cuony and Duncan Alexander, EPFL; details can
be found in Cuony et al., (2011). Right: spectral response of micromorph tandem cells with
no IRL, a 40 nm thick IRL and a 100 nm thick IRL: increasing the IRL thickness results in
enhanced (reduced) response in the 500–800 nm range in the top (bottom) cell. Courtesy of
Peter Cuony, IMT PV Lab, Neuchâtel.
Das et al., 2008). The best materials were shown to have a mixed-phase structure
with nanometre-wide silicon filaments embedded in an amorphous silicon oxide
matrix. This advanced nanostructure can be imaged by energy-filtered transmission
electron microscopy (EFTEM), as shown in Figure 4.20 (Cuony et al., 2012).
Light regions correspond to the doped silicon-rich phase that conducts, whereas
dark regions represent the silicon oxide phase that produces the material’s low
refractive index.
For optimum light harvesting in the top cell, the IRL must be combined
with efficient light scattering at the cell entrance, or it must be structured to
provide a textured interface to the incident photons, as shown in Figure 4.10
for a tandem device in the substrate configuration with an asymmetric reflector
(Söderström et al., 2010). In the superstrate configuration, i sc in 250 nm thick
top cells are typically about 11 mA cm−2 , while the integration of an IRL can
increase i sc up to 14 mA cm−2 , demonstrating the impact that such a thin layer can
have on efficiency. While n-type silicon oxide nanomaterials are straightforward
IRL solutions adaptable to mass production, consecutive approaches such as dis-
tributed Bragg reflectors and more sophisticated structures such as 3D photonic
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch04 page 176
crystals are also under development at the cell level to further control the spectral
selectivity and the light distribution (Bielawny et al., 2008). Nano-patterned low
refractive index dielectric materials (n r ∼ 1.3–1.5 at 600 nm) were also demon-
strated to have a high potential as efficient IRL (Boccard et al., 2013), with in
that latter case even a potential for altering the roughness of the surface on which
grows the µc-Si:H bottom cell, so as to provide in addition a smoothening of the
surface.
reduction of red and NIR light transmission. High carrier mobility is essential to
ensure low sheet resistances (Rsheet ) for a TCO with reduced free carrier density.
Recent advances in the synthesis of doped ZnO have yielded mobilities of 40–
60 cm2 /Vs and absorption levels below 5% over the full spectral range for a layer
with Rsheet = 10 /sq, and around 2% for Rsheet = 20 /sq (Ruske et al., 2010;
Boccard et al., 2010).
Alternatively, hydrogen-doped indium oxide (In2 O3 :H) was recently reported
as a highly transparent TCO; even NIR absorption losses are low (<1%) thanks to a
high mobility of up to 120 cm2/Vs (Koida et al., 2010). These layers do not provide
a native or post-process texture, but can be used in next-generation superstrates as
thin and highly transparent TCO layers.
Figure 4.21 Left: SEM cross-section of a-Si:H/µc-Si:H tandem solar cell, developed on
a glass superstrate implementing a nanoimprinted UV lacquer that provides large features,
a thin In2 O3 :H layer allowing for conduction and high transparency, and a low-doped LP-
CVD ZnO layer adding small pyramidal features. Courtesy of Mathieu Boccard; details
in Boccard et al. (2012). Right: tilted SEM top view image of this multi-scale textured
superstrate. Courtesy of Mathieu Boccard, IMT PV Lab, Neuchâtel.
2011b). These recent developments are important tools to determine and imple-
ment the most promising light-trapping schemes on the glass or in the electrode
itself. Some examples of periodic to random structures transferred to the electrode
or replicated on the glass substrate are shown in Figure 4.22.
Multi-scale textures are particularly interesting approaches for tandem cells
since they can combine efficient light scattering by small and large textures for
the amorphous and microcrystalline cells, respectively. Care must be taken to keep
morphologies that are adapted for the growth of the µc-Si:H cell. Two-step etch-
ing procedures on sputtered ZnO result in a surface morphology combining large
craters with smaller features (Hüpkes et al., 2010). Also, multi-scale texturing
of CVD-grown TCO (ZnO and SnO2 ) was demonstrated (Taneda et al., 2007).
A complementary approach is the combination of textured TCO with textured
glass (Bailat et al., 2010; Kroll et al., 2011) or with nanoimprinted lacquer on
glass (Boccard et al., 2012; Bessonov, 2011).
Recently, micromorph cells with record stabilised efficiencies of 11.9% (Bailat
et al., 2010) and 12.3% (Kroll et al., 2011) were demonstrated by combining
textured glass with textured ZnO. Additionally, superstrates combining both multi-
scale textures and bilayer TCOs are under development, and Figure 4.21 shows an
a-Si:H/µc-Si:H tandem cell realised on such a substrate. Superstrates with further
combinations of multi-layer TCOs and multiple textures produced in either TCO or
glass by direct texturing or nanoimprinting are still under study to further enhance
the efficiency of thin-film silicon solar cells.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch04 page 179
Figure 4.22 SEM images of (a) as-grown random pyramidal LPCVD ZnO; (b) thermally
roughened silver; and (c) alkaline-etched crystalline silicon, all realised by nanoimprinting
in a transparent lacquer. Courtesy of Jordi Escarre and Karin Söderström, IMT PV Lab,
Neuchâtel; details in Söderström et al. (2010). SEM images of (d) as-grown thick LP-CVD
ZnO; (e) a one-dimensional periodic grating; and (f) quasi-periodic hexagonal dimple pattern
of anodic oxidation of aluminium, all transferred to the surface of ZnO front electrodes via
nanomoulding. Courtesy of Corsin Battaglia; details in Battaglia et al. (2011b). (g): a-Si:H
solar cell on a nanomoulded ZnO front electrode with a surface texture corresponding to (f).
et al., 2011; Bugnon et al., 2011). For instance, the implementation of silicon-rich
silicon oxide-doped layers was demonstrated to enhance the electrical performance
of thin-film silicon solar cells grown on highly textured substrates. These doped
layers were shown to mitigate the impact of localised defects on cell electrical
performance (Despeisse et al., 2011). Section 4.5 discussed how such defective
areas arise when silicon is deposited on highly textured surfaces exhibiting sharp
valleys (Python et al., 2008).
Moreover, plasma process conditions during microcrystalline silicon growth
were also shown to affect the quality and density of these localised defect regions
that occur over sharp underlying surface features (Bugnon et al., 2011), demonstrat-
ing that globally dense µc-Si:H material has to be achieved on textured substrates
to maximise efficiency. Thin devices were shown to have high potential (Schicho
et al., 2010), and the enhanced resilience of the solar cells to the substrate texture
allowed for the development of high-efficiency thin superstrate devices, with 11.3%
and 10.9% stable cells realised using 1.1 µm and 0.8 µm thick bottom cells, respec-
tively (Despeisse et al., 2011; Bailat et al., 2010). Such thin cells give stabilised
efficiencies higher than 10% in modules (Kratzla et al., 2010), and they trigger cost
reductions because of reduced deposition and reactor cleaning times, resulting in
a promising trade-off between efficiency and costs for the micromorph approach.
Alternatively, high-efficiency concepts with folded 3D cell designs (Vanecek
et al., 2010; Naughton et al., 2010) are under consideration for a-Si:H single junc-
tions and micromorph tandem devices. Novel nanostructures such as nano-columns
or nano-/micro-holes on superstrates or substrates could open possibilities; the cell
layers would be optically thick but electrically thin (cf. Figure 4.23). Accordingly,
the resulting folded cell would yield thin layers with very low LID (Vanecek et al.,
2010).
Figure 4.23 SEM image of aligned nano-pillars (left panel (Rizal et al., 2013)) and FIB
cross-section through n-i- p solar cells deposited on these pillars. The inset illustrates non-
conformal growth with preferential deposition of reflector, silicon and ITO layer at the tips
(Naughton et al., 2010).
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch04 page 181
on the requirement to ‘decouple’ the electrical from the optical transport paths, a
simple calculation based on the best top and bottom cell realised so far shows that
tandem solar cells are probably limited to stabilised efficiencies between 14% and
15%. For two reasons, triple-junction solar cells such as a-Si:H/µc-Si:H/µc-Si:H
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch04 page 183
Table 4.3 Overview of noticeable efficiencies (reported by the authors, not certified).
liquid crystal displays (LCDs); the TFTs in these displays are based on amorphous
silicon as well as silicon nitride. Moreover, the FPD industry demonstrated a con-
tinuous decrease of display production costs (20% per m2 annually). It was hence
not a surprise that equipment makers, who had made a significant investment in
the development of coating tools, identified thin-film silicon solar technology as an
emerging growth market for their products, with a focus on PECVD equipment that
could be used for layer deposition. This led companies such as Oerlikon, Applied
Films (later purchased by Applied Material), Ulvac and Leybold Optics to enter
the PV market, trying to transfer several of the processes developed in research
laboratories to mass production. Most manufacturers work on Gen5 scale (1.4 m2)
but Applied Materials has also delivered tools for Gen8.5 (5.5 m2). The transition
from a mobility-controlled device (as in a TFT) to a lifetime-controlled device (as
in PV) was, however, not trivial for three related reasons:
• In a-Si:H and µc-Si:H, the material quality depends very much on the deposi-
tion conditions and rate. Reasonable rates at acceptable quality often required
equipment adaptation (e.g. narrower inter-electrode gaps for higher pressure
or measures to avoid standing-wave patterns for the case of VHF-excitation).
• By definition, µc-Si:H is a material deposited close to a phase transition; con-
sequently its homogeneous deposition on >1 m2 required significant hardware
adaptation in order to compensate for gas flow and electromagnetic inhomo-
geneities, as well as boundary effects.
• The two first steps had to be taken in conjunction with the development of
the TCO front contact (not fully developed at that time), the back contact and
laser scribing.
While solving these issues, several equipment makers were able to sign con-
tracts to sell ‘turnkey’ production lines, making thin-film silicon the first technology
offering an easy entrance into the PV market, while covering the full chain (from
glass to module). Indeed from 2006 to 2008, concerned about silicon feedstock
problems for c-Si cells manufacturers, over 50 companies announced their entrance
into thin-film silicon module manufacturing. Even though solving the three issues
noted above was more demanding than initially expected, in 2011 several com-
panies were able to manufacture modules with total-area stabilised efficiencies of
6.5–7.3% and 9–10% for a-Si:H and a-Si:H/µc-Si:H modules on glass, respec-
tively. Note that ‘total-area’ refers to a widely used module size of 1.4 m2 which
includes a non-active edge area of 6–8%. Commercial a-Si:H/a-Si1−x Gex :H/a-
Si1−y Ge y :H triple-junction modules on flexible substrates reach 8.2% stabilised
efficiencies (designated aperture area); a comparison with the record devices in
Table 4.2 shows that there is a relatively little gap between small-scale cell results
and m2 sized modules, at least for p-i -n cells.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch04 page 185
Table 4.4 Overview of advantages and challenges linked to the use of various front TCOs
in amorphous and a-Si:H/µc-Si:H production.
and the TCO. Typical TCO thickness is 700 nm. After SnO2 :F deposition, the
glass is cut and the edges are ground.
• Off-line deposition of SnO2 :F by APCVD: deposition on glass plates that are
reheated to typical temperatures of 540–590◦C. Deposition rates are 50 nm/s
and the same precursors can be used as for the inline process. Typical TCO
thickness is 700 nm.
• Off-line sputter-etched ZnO: aluminium-doped ZnO is sputtered and subse-
quently wet-etched to create crater-like nanostructures (Kluth et al., 2003).
ZnO rotary ceramic targets containing 1–2 wt.% of Al2 O3 are used for DC
sputtering (Zhu et al., 2009) at typical temperatures of 300–370◦C. After depo-
sition, the cooled substrates are etched in a 0.3–0.5% HCl solution. Typical
TCO thickness is 700 nm.
• Off-line deposition of ZnO by LPCVD (Vogler et al., 2008): deposition typi-
cally takes place at 170–200◦C, at pressures of a few mbar, using di-ethyl-zinc
(DEZ, (C2 H5 )2 Zn), water and di-borane for doping. Typical deposition rates
are 2–4 nm per second, and typical layer thickness is 1.4–2 µm.
Notably, the processes run by research laboratories might differ from those
employed by industry. At the laboratory level, all of these TCOs have demonstrated
good results. Between 2009 and 2011, certified world record efficiencies were
reported for amorphous cells deposited on LP-CVD ZnO (Benagli et al., 2009)
and for micromorph cells on LP-CVD ZnO grown on textured glass (Bailat et al.,
2010). Micromorph modules with efficiencies of over 10% have been reported
for most of the above TCOs (Klein et al., 2010; Kratzla et al., 2010; Aya et al.,
2011).
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch04 page 187
and processing for µc-Si:H with solutions based on parallel plate RF PECVD.
In most approaches, the inter-electrode gap is low (8–20 mm) and the pressure is
kept high (5–15 mbars); under these conditions, collisions slow down ions that are
accelerated in the high voltage drop at the sheath. Most deposition regimes use
high dilution (at most a few percent of silane in hydrogen). In very large coating
systems (such as Gen8.5 tools), a combination of multiple small hollow cathodes
with variable geometry may be integrated in the RF electrode and combined with a
gas diffuser consisting of holes in the showerhead plate. Additionally varying the
spacing or the gas conductance is varied, such designs can compensate standing
wave and edge effects.
A wide variety of system configurations exists (e.g. inline, vertical, batch,
cluster). TEL (formerly Oerlikon) Solar uses the design shown in Figure 4.24 where
Figure 4.24 Top: Example of a compact production system (KAI systems for PECVD,
with two process modules with stacks of 10 reactors). Bottom: Flexible amorphous silicon
modules for roof-top integration.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch04 page 189
a stack of ten reactors is placed inside a single vacuum chamber which reduces the
required components (common gas supply, pressure control, and vacuum pump
for all reactors). Only the 40.68 MHz VHF excitation is supplied from individual
power supplies. The standing wave effect is suppressed using a magnetic lens
compensation system (Sansonnens and Schmitt, 2003). Several other equipment
manufacturers, some of them not from the FPD industry, also proposed low-cost
systems for amorphous silicon, usually based on 13.56 MHz parallel plate reactors
and sometimes with lateral ingress of the processing gas (e.g. with one RF generator
powering one electrode and with two counter electrodes, allowing for coating of
four glass substrates with one generator).
Most of today’s equipment is suitable for deposition rates of 0.2–0.4 nm/s for
a-Si:H and 0.3–0.6 nm/ for µc-Si. Approaches for much higher deposition on large-
area cells use arrays of nozzle electrodes and local pumping (Aya et al., 2011) which
probably prevents excessive powder formation and plasma perturbation, or VHF at
60 MHz with ladder-shaped electrodes (Takeuchi et al. 2001; Takatsuka et al., 2005;
Mashima et al., 2006). This last configuration creates an asymmetry favouring low-
energy bombardment of the substrate. A dual frequency mode and duty cycle with
suitable spacing of the ladder allows compensating the standing wave issue. These
last two approaches have facilitated good initial module efficiencies over 10%
at high deposition rate (>2 nm/s). However, for such systems, input power and
thermal load management are not trivial.
at high temperature, aluminium and particularly silver develop surface textures that
are suitable for light trapping (Banerjee and Guha, 1991). This ‘hot silver’ reflector
is also applicable to polyimide substrates, but low-cost plastic such as polyethylene
does not withstand the required temperatures. Alternatively, flat silver covered
by textured ZnO can act as a high-quality reflector (Guha et al., 2011), or light
trapping textures may be embossed into the plastic substrate itself (Bailat et al.,
2005; Escarré et al., 2011).
Table 4.5 summarises the status of some flexible thin-film silicon companies.
Usually, n-i - p modules incorporate EVA/ETFE protection at the front side, which
is not vapour tight but is compatible with the front ITO used by most manufac-
turers. Contrary to the case of glass substrates, there are no official turnkey sup-
pliers for roll-to-roll technology and most developments have been made in-house
by the respective companies. Helianthos uses the p-i -n structure on a sacrificial
aluminium substrate coated with a SnO2 front contact, a substrate that is also
compatible with micromorph technology (van den Donker et al., 2007).
Table 4.5 Overview of some flexible thin-film silicon suppliers and developers.
Module Capacity
Company Country Configuration efficiency (2011) Notes
lower than those given for thicker devices (Fthenakis et al., 2008a; Fthenakis et al.,
2008b; Bravi et al., 2011).
Acknowledgements
We would like to thank Z. Holman for careful reading and corrections, D. Sheel
(CVtech) and J. Hüpkes (HZB Jülich) for providing information on TCO prepa-
ration, Reinhardt Benz and Irene Steimen (Oerlikon Solar) for providing cost cal-
culation data, the personnel of PV Lab for providing figures and data, as well as
D. Alexander and M. Cantoni (CIME EPFL) and M. Leboeuf (CSEM) for creating
some of the TEM and SEM images.
References
Aya Y., Shinohara W., Matsumoto M., Murata K., Kunii T., Nakagawa M., Terakawa A. and
Tanaka M. (2011), ‘Progress of thin-film silicon photovoltaic technologies SANYO’,
Progr. Photovoltaics: Res. Applications, published online.
Bae S., Kaan Kalkan A., Cheng S. and Fonash S. (1998), ‘Characteristics of amorphous
and polycrystalline silicon films deposited at 120◦ C by electron cyclotron resonance
plasma-enhanced chemical vapor deposition’, J. Vac. Sci. Technol. A 16, 3, 1992.
Bailat J., Dominé D., Schlüchter R, Steinhauser J., Faÿ S., Freitas F., Bucher C., Feitknecht
L., Niquille X., Tscharner R., Shah A. and Ballif C. (2006), ‘High efficiency pin
microcrystalline and micromorph thin-film silicon solar cells deposited on LPCVD
ZnO coated glass substrates’, Proc. 4th World PVSEC, Honolulu, HI.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch04 page 194
Bailat J., Fesquet L., Orhan J.-B., Djeridane Y., Wolf B., Madliger P., Steinhauser J., Benagli,
Borrello D., Castens L., Monteduro G., Dehbozorgi B., Vallat-Sauvain E., Multone X.,
Romang D., Boucher J.-F., Meier J., Kroll U., Despeisse M., Bugnon G., Ballif C.,
Marjanovic S., Konhke G., Borrelli N., Koch K., Liu J., Modavis R., Thelen D., Vallon
S., Zakharian A. and Weidman, D. (2010), ‘Recent developments of high-efficiency
micromorph tandem solar cells in KAI-M PECVD reactors’,Proc. 5th World PVSEC,
Valencia, pp. 2720–2724, 3BO.11.5.
Bailat J., Terrazzoni-Daudrix V., Guillet J., Freitas F., Niquille X., Shah A., Ballif C., Scharf
T., Morf R., Hansen A., Fischer D., Ziegler Y. and Closset A. (2005), ‘Recent develop-
ment of solar cells on low-cost plastic substrates’, Proc. 20th EU Photovoltaic Solar
Energy Conf., Barcelona, pp. 1529–1532.
Ballif C., Barraud L., Battaglia C., Boccard M., Bugnon G., Charrière M., Cuony P.,
Despeisse M., Ding L., Escarré J., Haug F.-J., Hänni S., Löfgren L., Sculatti-Meillaud
S., Nicolay S., Parascondolo G., De Wolf S., Söderström K. and Stückelberger M.
(2011), ‘Novel Materials and Superstrates for High-Efficiency Micromorph Solar
Cells’, Proc. 26th EUPVSC, Hamburg, pp. 2384–2391, 3CO.1.1.
Ballutaud J., Bucher C., Hollenstein Ch., Howling A. A., Kroll U., Benagli S., Shah A. and
Buechel A. (2004), ‘Reduction of the boron cross-contamination for plasma deposition
of p-i-n devices in a single-chamber large area radio-frequency reactor’, Thin Solid
Films 468, 222–225.
Banerjee A., Beglau D., Su T., Pietka G., Yue G., Yan B., Yang J. and Guha S. (2011a), ‘11.0%
stable efficiency on large area, encapsulated a-Si:H and a-SiGe:H based multijunction
solar cells using VHF technology’, Proc. MRS Spring Meeting, San Francisco, CA,
A14-02.
Banerjee A. and Guha S. (1991), ‘Study of back reflectors for amorphous-silicon alloy
solar-cell application’, J. Appl. Phys. 69, 1030–1035.
Banerjee A., Liu F., Beglau D., Su T., Pietka G., Yang J. and Guha S. (2011c), ‘12.0%
efficiency on large-area, encapsulated, multijunction nc-Si:H triple cell’, Proc. 37th.
IEEE PVSC, Seattle, WA, pp. 1013–1017.
Banerjee A., Su T., Beglau D., Pietka G., Liu F., Yan B., Yang J. and Guha S. (2011b), ‘High
efficiency, large area, nanocrystalline silicon based triple junction solar cells’, Proc.
MRS Spring Meeting, San Francisco, CA, A01-03.
Bano G., Horvath P., Rozsa K. and Gallagher A. (2005), ‘The role of higher silanes in
silane-discharge particle growth’, J. Appl. Phys. 98, 013304.
Bartlomé R., Strahm B., Sinquin Y., Feltrin A. and Ballif C. (2010), ‘Laser applications in
thin-film photovoltaics’, Appl. Phys. B 100, 427–436.
Basore P.A. (2002), ‘Pilot production of thin-film crystalline silicon on glass modules’,
Proc. 29th IEEE PVSC, New Orleans, LA, pp. 49–52.
Battaglia C., Escarre J., Söderström K., Erni L., Ding L., Bugnon G., Billet A., Boccard M.,
Barraud L., De Wolf S., Haug F.-J., Despeisse M. and Ballif C. (2011a), ‘Nanoimprint
lithography for high efficiency thin-film silicon solar cells’, Nano Letters 11, 661–665.
Battaglia C., Escarre J., Söderström K., Charrière M., Despeisse M., Haug F.-J. and Ballif C.
(2011b), ‘Nanomoulding of transparent zinc oxide electrodes for efficient light trapping
in solar cells’, Nature Photonics, DOI 10.10038.
Benagli S., Borrello D., Vallat-Sauvain E., Meier J., Kroll U., Hoetzel J., Bailat J.,
Steinhauser J., Marmelo M., Monteduro G. and Castens L. (2009), ‘High efficiency
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch04 page 195
Buehlmann P., Bailat J., Domine D., Billet A., Meillaud F., Feltrin A. and Ballif C. (2007),
‘In situ silicon oxide based intermediate reflector for thin-film silicon micromorph solar
cells’, Appl. Phys. Lett. 91, 143505.
Bugnon G., Parascandolo G., Söderström T., Bartlome R., Cuony P., Hänni S., Boccard M.,
Holovsky J., Despeisse M., Meillaud F. and Ballif C. (2011), ‘High rate deposition of
microcrystalline silicon with silicon oxide doped layers: highlighting the competing
roles of both intrinsic and extrinsic defects on the cells performances’, to be published
in Proc. 37th IEEE PVSC, Seattle, WA.
Campbell P. and Green M. A. (1987), ‘Light trapping properties of pyramidally textured
surfaces’, J. Appl. Phys. 62, 243–249.
Carlson D. E. and Wronski C. R. (1976), ‘Amorphous silicon solar cells’, Appl. Phys. Lett.
28, 671–673.
Carniglia C. K. (1979), ‘Scalar scattering theory for multilayer optical coatings’, Optical
Engineering 18, 104–115.
Chen T., Huang Y. L., Yang D., Carius R. and Finger F. (2010), ‘Microcrystalline silicon
thin-film solar cells with microcrystalline silicon carbide window layers and silicon
absorber layers both prepared by hot-wire CVD’, Physica Stat. Solidi, Rapid Research
Lett. 4, 61–63.
Chen T., Huang Y., Yang D., Carius R. and Finger F. (2011), ‘Development of micro-
crystalline silicon carbide window layers by hot-wire CVD and their applications in
microcrystalline silicon thin-film solar cells’, Thin Solid Films 519, 4523–4526.
Chittick R., Alexander J. and Sterling H. (1969), ‘The preparation and properties of amor-
phous silicon’, J. Electrochem. Soc. 116, 77–81.
Collins R., Ferlauto A., Ferreira G., Chen C., Koh J., Koval R., Lee Y., Pearce J. and Wronski
C. (2003), ‘Evolution of microstructure and phase in amorphous, protocrystalline, and
microcrystalline silicon studied by real time spectroscopic ellipsometry’, Solar Energy
Mater. Solar Cells 78, 143–180.
Cuony P., Alexander D., Perez-Wurfl Y., Despeisse M., Bugnon G., Boccard M., Söderström
T., Hessler-Wyser H., Hébert C. and Ballif C. (2012), ‘Silicon filaments in silicon oxide
for next generation photovoltaics’, Adv. Mater. 24, 1182–1186.
Cuony P., Marending M., Alexander D. T. L., Boccard M., Bugnon G., Despeisse M. and
Ballif C. (2010), ‘Mixed-phase p-type silicon oxide containing silicon nanocrystals
and its role in thin-film silicon solar cells’, Appl. Phys. Lett. 97, 213502-1–21502-3.
Curtins H., Favre M., Wyrsch N., Brechet M., Prasad K. and Shah A. (1987), ‘High-rate
deposition of hydrogenated amorphous silicon by the VHF-GD method’, Proc. 19th
IEEE PVSC, New Orleans, LA, pp. 695–698.
Das C., Lambertz A., Huepkes J., Reetz W. and Finger F. (2008), ‘A constructive combination
of antireflection and intermediate-reflector layers for a-Si/µc-Si thin-film solar cells’,
Appl. Phys. Lett. 92, 053509-1–053509-3.
Deckman H. W., Wronski C. R., Witzke H. and Yablonovitch E. (1983), ‘Optically enhanced
amorphous-silicon solar-cells’, Appl. Phys. Lett. 42, 968–970.
Delli P., Veneri L. Mercaldo V. and Usatii L. (2010), ‘Silicon oxide based n-doped layer for
improved performance of thin-film silicon solar cells’, Appl. Phys. Lett. 97, 023512–
023515.
Despeisse M., Battaglia C., Boccard M., Bugnon G., Charrière M., Cuony P., Hänni
S., Löfgren L., Meillaud F., Parascandolo G., Söderström T. and Ballif C. (2011),
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch04 page 197
conductive oxide/Si interface by TiO2 /ZnO antireflection coating’, Appl. Phys. Lett.
88, 183508-1–183508-3.
Gee J. M. (2002), ‘Optically enhanced absorption in thin silicon layers using photonic
crystals’, Proc. 29th IEEE PVSC, New Orleans, LA, pp. 150–153.
Green M. A., Basore P., Chang N., Clugston D., Egan R., Evans R., Hogg D., Jarnason S.,
Keevers M. and Lasswell P. (2004), ‘Crystalline silicon on glass (CSG) thin-film solar
cell modules’, Solar Energy 77, 857–863.
Guha S., Narasimhan K. and Pietruszko S. (1981), ‘On light induced effect in amorphous
hydrogenated silicon’, J. Appl. Phys. 52, 859–860.
Guha S., Yang J. and Banerjee A. (2000), ‘Amorphous silicon alloy photovoltaic research —
present and future’, Progr. Photovoltaics: Res. Applications 8, 141–150.
Guha S., Yang J., Pawlikiewicz A., Glatfelter T., Ross R. and Ovshinsky S. (1989), ‘Band
gap profiling for improving the efficiency of amorphous silicon alloy solar cells’, Appl.
Phys. Lett. 54, 2330–2332.
Guha S., Yang J. and Yan B. (2011), ‘Amorphous and nanocrystalline silicon solar cells
and modules’, in Bhattacharya, P., Fornari, R. and Kamimura, H. (eds) Comprehensive
Semiconductor Science and Technology, Elsevier, London.
Haas S., Gordijn A. and Stiebig H. (2007), ‘High speed laser processing for monolithical
series connection of silicon thin-film modules’, Progr. Photovoltaics: Res. Applications
16, 195–203.
Hanak J. J. and Korsun V. (1982), ‘Optical stability studies of a-Si: H solar cells’, Proc.
16th IEEE PVSC, San Diego, CA, pp. 1381–1383.
Haneman D. (1968), ‘Electron paramagnetic resonance from clean single-crystal cleavage
surfaces of silicon’, Phys. Rev. 170, 705–718.
Hänni S., Battaglia C., Boccard M., Bugnon G., Cuony P., Despeisse M., Nicolay S., Meil-
laud F. and Ballif C. (2011), ‘Towards better understanding of long-term stability in
thin-film microcrystalline solar cells’, Proc. 26th PVSEC, Hamburg, pp. 2699–2703,
3AV.2.42.
Haug F.-J., Söderström K., Naqavi A. and Ballif C. (2011), ‘Resonances and absorp-
tion enhancement in thin-film silicon solar cells’, J. Appl. Phys. 109, 084516-1–
084516-8.
Haug, F.-J. Söderström, T., Python, M., Terrazzoni-Daudrix, V., Niquille, X. and Ballif, C.
(2009), ‘Development of micromorph tandem solar cells on flexible low cost plastic
substrates’, Solar Energy Mater. Solar Cells 93, 884–8877.
Higashi G., Chabal Y., Trucks G. and Raghavachari K. (1990), ‘Ideal hydrogen termination
of the Si (111) surface’, Appl. Phys. Lett. 56, 656–658.
Hirasaka M., Suzuki K., Nakatani K., Asano M., Yano M. and Okaniwa H. (1990), ‘Design
of textured Al electrode for a hydrogenated amorphous silicon solar cell’, Solar Energy
Mater. 20, 99–110.
Hüpkes J., Owen J. I., Bunte E., Zhu H., Pust S. E., Worbs J. and Jost G. (2010), ‘New
texture etching of zinc oxide: tunable light trapping for silicon thin-film solar cells’,
Proc. 5th World PVSEC, Valencia, pp. 3224–3227, 3AV.2.25.
Inthisang S., Krajangsang T., Yunaz I. A., Yamada A., Konagai M. and Wronski C. (2011),
‘Fabrication of high open circuit voltage a-Si1−x O x :H solar cells by using p-a-
Si1−x Ox :H as window layer’, Physica Stat. Solidi (c), 2990–2993.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch04 page 199
Kratzla T., Zindel A. and Benz R. (2010), ‘Oerlikon Solar’s key performance drivers to grid
parity’, Proc. 5th. World PVSEC, Valencia, pp. 2807–2810, 3CO.14.1.
Krč J., Smole F. and Topič M. (2003), ‘Analysis of light scattering in amorphous Si:H
solar cells by a one-dimensional semi-coherent optical model’, Progr. Photovoltaics
11, 15–26.
Kroll U., Meier J., Torres P., Pohl J. and Shah A. (1998), ‘From amorphous to micro-
crystalline silicon films prepared by hydrogen dilution using the VHF (70 MHz) GD
technique’, J. Non-Cryst. Solids 227–230, 68–72.
Kroll U., Meier J., Steinhauser J., Fesquet L., Benagli S., Orhan J.-B., Djeridane Y., Wolf
B., Borrello D., Madliger P., Vallat-Sauvain E., Boucher J.-F., Castens L., Multone X.,
Marjanovic S., Konhke G., Borrelli N., Koch K., Liu J., Modavis R., Thelen D., Val-
lon S., Zakharian A. and Weidman D. (2011), ‘Recent developments of high-efficiency
micromorph tandem solar cells in KAI-M PECVD reactors’, Proc. 26th. PVSEC, Ham-
burg, pp. 2340–2343, 3BO.2.6.
Kroll U., Shah A., Keppner H., Meier J., Torres P. and Fischer D. (1997), ‘Potential of VHF
plasmas for low-cost production of a-Si:H solar cells’, Solar Energy Mater. Solar Cells
48, 343–350.
Laaziri K., Kycia S., Roorda S., Chicoine M., Robertson J., Wang J. and Moss S. (1999),
‘High-energy x-ray diffraction study of pure amorphous silicon’, Phys. Rev. B60,
13520–13533.
Lambertz A., Dasgupta A., Reetz W., Gordijn A., Carius R. and Finger F. (2007), ‘Micro-
crystalline silicon oxide as intermediate reflector for thin-film silicon’, Proc. 22nd
EU-PVSEC, Milan, pp. 1838–1842.
Lechner P., Frammelsberger W., Psyk W., Geyer R., Maurus H., Lundszien D., Wagner H.
and Eichhorn B. (2008), ‘Status of performance of thin-film siliocn solar cells and
modules’, Proc. 23rd EU-PVSEC, Valencia, pp. 2023–2026.
Lechner P., Geyer R., Haslauer A. and Roehrl T. (2010), ‘Long-term performance of ASI
tandem junction thin-film solar modules’, Proc. 5th. World PVSEC, Valencia, pp. 3283–
3287, 3AV.2.46.
Li H., Franken R. H., Stolk R. L., Rath J. K. and Schropp R. E. I. (2008), ‘Mechanism of
shunting of nanocrystalline silicon solar cells deposited on rough Ag/ZnO substrates’,
Solid State Phenomena 131/133, 27–32.
Li Y.-M., Jackson F., Yang L., Fieselmann B. F. and Russel L. (1994), ‘An exploratory
survey of p-layers for a-Si:H solar cells’, Proc. MRS Spring Meeting, San Francisco,
CA, pp. 663–668.
Lucovsky G. (1992), ‘Hydrogen in amorphous silicon: local bonding and vibrational prop-
erties’, J. Non-Cryst. Solids 141, 241–256.
Lundszien D., Finger F. and Wagner H. (2002), ‘Band-gap profiling in amorphous silicon-
germanium solar cells’, Appl. Phys. Lett. 80, 1655.
Mahan A. H., Williamson D. L., Nelson B. P. and Crandall R. S. (1989), ‘Small angle X-ray
scattering studies of microvoids in a-SiC:H and a-Si:H’, Solar Cells 27, 465–476.
Mahan A. H., Xu Y., Iwaniczko E., Williamson D. L., Nelson B. P. and Wang Q. (2002),
‘Amorphous silicon films and solar cells deposited by HWCVD at ultra-high deposition
rates’, J. Non-Cryst. Solids, 299–302, 2–8.
Mai Y., Klein S., Carius R., Stiebig H., Geng X. and Finger F. (2005), ‘Open circuit voltage
improvement of high-deposition-rate microcrystalline silicon solar cells by hot wire
interface layers’, Appl. Phys. Lett. 87, 073503-1–073503-12.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch04 page 201
Mashima H., Yamakoshi H., Kawamura K., Takeuchi Y., Noda M., Yonekura Y., Takatsuka
H., Uchino S. and Kawai Y. (2006), ‘Large area VHF plasma production using a ladder-
shaped electrode’, Thin Solid Films 506–507, 512–516.
Matsuda A. (1983), ‘Formation kinetics and control of microcrystallite in µc-Si: H from
glow discharge plasma’, J. of Non-Crystalline Solids 59, 767–774.
Matsuda A. (1999), ‘Growth mechanism of microcrystalline silicon obtained from reactive
plasmas’, Thin Solid Films 337, 1–6.
Matsuda A. and Ganguly G. (1995), ‘Improvement of hydrogenated amorphous silicon
germanium alloys using low power disilane–germane discharges without hydrogen
dilution’, Appl. Phys. Lett. 67, 1274–1276.
Matsuda A., Koyama M., Ikuchi N., Imanishi Y. and Tanaka K. (1986), ‘Guiding principle
in the preparation of high-photosensitive hydrogenated amorphous Si-Ge alloys from
glow-discharge plasma’, Japan. J. Appl. Phys. 25, L54–L56.
Matsuda A., Yamasaki S., Nakagawa K., Okushi H., Tanaka K., Iizima S., Matsumura M.
and Yamamoto H. (1980), ‘Electrical and structural properties of phosphorous-doped
glow-discharge Si:F and Si:H films’, Japanese J. Appl. Phys. 19, L305–L308.
Matsui T., Jia H. and Kondo M. (2010), ‘Thin-film solar cells incorporating microcrystalline
Si1−x Gex as efficient infrared absorber: an application to double junction tandem solar
cell’, Progr. Photovoltaics 18, 48–53.
Matsui T., Kondo M., Nasuno Y., Sonobe H. and Shimizu S. (2006), ‘Key issues for fab-
rication of high quality amorphous and microcrystalline solar cells’, Thin Solid Films
501, 243–246.
Matsui T., Sai H. and Kondo. (2013), ‘Development of highly stable and efficient amorphous
silicon based solar cells’, Proc. 28th European PV Conf., Paris, p. 3DO.7.2.
McCurdy R. J. (1999), ‘Successful implementation methods of atmospheric CVD on a glass
manufacturing line’, Thin Solid Films 351, 66–72.
Meier J., Dubail S., Flückinger R., Fischer D., Keppner H. and Shah A. (1994), ‘Intrinsic
microcrystalline silicon — a promising new thin-film solar cell material’, Proc. 1st
World PVSEC, Honolulu, HI, pp. 409–412.
Meier J., Spitznagel J., Kroll U., Bucher C., Faÿ S., Moriarty T. and Shah A. (2003), ‘High
efficiency amorphous and micromorph silicon solar cells’, Proc. 3rd World PVSEC,
Osaka, pp. 2801–2805.
Meier J., Torres P., Platz R., Dubail S., Kroll U., Selvan J. A., Pellaton Vaucher N., Hof
C., Fischer D., Keppner H. and Shah A. (1996), ‘On the way towards high-efficiency
thin-film silicon solar cells by the ‘micromorph’ concept’, Proc. MRS Spring Meeting,
420, 3–14.
Nakamura G., Sato K., Yukimoto Y. and Shirahata K. (1980), ‘Amorphous Si1−x Gex for
high performance solar cell’, Proc. 3rd EU-PVSEC, Cannes, pp. 835–839.
Nakano S., Matsuoka T., Kiyama S., Kawata H., Nakamura N., Nakashima Y., Tsuda
S., Nishiwak H., Ohnishi M., Nagaoka I. and Kuwano Y. (1986), ‘Laser patterning
method for integrated type a-Si solar cell submodules’, Japan. J. Appl. Phys. 25,
1936–1943.
Nasuno Y., Kondo M. and Matsuda A. (2001a), ‘Effects of substrate surface morphology
on microcrystalline silicon solar cells’, Japan. J. Appl. Phys. 40, L303–L305.
Nasuno Y., Kondo M. and Matsuda A. (2001b), ‘Passivation of oxygen-related donors
in microcrystalline silicon by low temperature deposition’, Appl. Phys. Lett. 78,
230–232.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch04 page 202
Naughton M. J., Kempa K., Ren Z. F., Gao Y., Rybczynski J., Argenti N., Gao W., Wang
Y., Peng Y., Naughton J. R., Burns M. J., Shepard A., Clary M., Ballif C., Haug F.-J.,
Söderström T., Cubero O. and Eminian C. (2010), ‘Efficient nano-coax based solar
cells’, Physica Stat. Solidi Rapid Research Lett. 4, 181–183.
Nishimoto T., Takai M., Miyahara H., Kondo M. and Matsuda A. (2002), ‘Amorphous silicon
solar cells deposited at high growth rate’, J. Non-cryst. Solids 299–302, 1116–1122.
Pieters B., Krč J. and Zeman M. (2007), ‘Advanced numerical simulation tool for solar
cells-ASA5’, Proc. 4th World PVSEC, Honolulu, HI, pp. 1513–1516.
Poruba A., Fejfar A., Remes Z., Springer J., Vanecek M., Kocka J., Meier J., Torres P. and
Shah A. (2000), ‘Optical absorption and light scattering in microcrystalline silicon
thin-films and solar cells’, J. Appl. Phys. 88, 148–160.
Powell M. and Deane S. (1996), ‘Defect-pool model and the hydrogen density of states in
hydrogenated amorphous silicon’, Phys. Rev. B 53, 10121–10132.
Python M., Dominé D., Söderström T., Meillaud F. and Ballif C. (2010), ‘Microcrystalline
silicon solar cells: effect of substrate temperature on cracks and their role in post
oxidation’, Progr. Photovoltaics 18, 491–499.
Python M., Vallat-Sauvain E., Bailat J., Dominé D., Fesquet L., Shah A. and Ballif C. (2008),
‘Relation between substrate surface morphology and microcrystalline silicon solar cell
performance’, J. Non-Cryst. Solids 354, 2258–2262.
Rech B., Repmann T., van den Donker M. N., Berginski M., Kilper T., Hupkes J., Calnan S.,
Stiebig H. and Wieder S. (2006), ‘Challenges in microcrystalline silicon based solar
cell technology’, Thin Solid Films 511, 548–555.
Redfield D. (1974), ‘Multiple pass thin-film silicon solar cell’, Appl. Phys. Lett. 25, 647–648.
Redfield D. (1982), ‘Energy-band tails and the optical absorption edge; the case of a-Si:H’,
Solid State Commun. 44, 1347–1349.
Ringbeck R. and Stutterlueti J. (2012), ‘BOS costs: status and optimization to reach industrial
grid parity’, Proc. 27th EU-PVSEC, Frankfurt, pp. 2961–2975, 4DP.2.2.
Rizal B., Ye F., Dhakal P., Chiles T.C., Shepard S., McMahon G., Burns M.J. and Naughton
M.J. (2013), ‘Imprint-templated nanocoax array architecture: Fabrication and utiliza-
tion’, Nano-Optics for Enhancing Light-Matter Interactions on a Molecular Scale,
NATO Science for Peace and Security Series B: Physics and Biophysics, 359–370.
Roca i Cabarrocas P. (2000), ‘Plasma enhanced chemical vapour deposition of amorphous,
polymorphous and microcrystalline silicon films’, J. Non-Cryst. Solids, 266–269,
31–37.
Rockstuhl C., Fahr S., Bittkau K., Beckers T., Carius R., Haug F.-J., Söderström T., Ballif
C. and Lederer F. (2011), ‘Comparison and optimization of randomly textured surfaces
in thin-film solar cells’, Optics Express 18, A335–A342.
Ruske F., Roczen M., Lee K., Wimmer M., Gall S., Hüpkes J., Hrunski D. and Rech B.
(2010), ‘Improved electrical transport in Al-doped zinc oxide by thermal treatment’,
J. Appl. Phys. 107, 013708–013708-8.
Sai H., Kanamori Y. and Kondo M. (2011), ‘Flattened light-scattering substrate in thin film
silicon solar cells for improved infrared response’, Appl. Phys. Lett. 98, 113502-1–
113502-3.
Sai H., Matsui T., Matsubara K., Kondo M. and Yoshida I. (2014), ‘11.0%-Efficient Thin
Film Microcrystalline Silicon Solar Cells with Honeycomb Textured Substrates’, pre-
sented at the 40th IEEE PVSC, Denver.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch04 page 203
Saito K., Sano M., Okabe S., Sugiyama S. and Ogawa K. (2005), ‘Microcrystalline silicon
solar cells fabricated by VHF plasma CVD method’, Solar Energy Mater. Solar Cells
86, 565–575.
Sakai H., Yoshida T., Hama T. and Ichikawa Y. (1990), ‘Effects of surface morphology of
transparent electrode on the open-circuit voltage in a-Si: H solar cells’, Japan. J. Appl.
Phys. 29, 630–635.
Sansonnens L., Schmidt H., Howling A. A., Hollenstein C., Ellert C. and Buechel A. (2006),
‘Application of the shaped electrode technique to a large area rectangular capacitively
coupled plasma reactor to suppress standing wave nonuniformity’, J. Vac. Sci. Technol.
A 24, 1425–1430.
Sansonnens L. and Schmitt J. (2003), ‘Shaped electrode and lens for a uniform radio-
frequency capacitive plasma’, Appl. Phys. Lett. 82, 182–184.
Schicho S., Hrunski D., van Aubel R. and Gordijn A. (2010), ‘High potential of thin (<1 µm)
a-Si:H/µc-Si:H tandem solar cells’, Progr. Photovoltaics 18, 83–89.
Schropp R. E. I., Carius R. and Beaucarne G. (2007), ‘Amorphous silicon, microcrystalline
silicon, and thin-film polycrystalline silicon solar cells’, Proc. Mater. Research Sym-
posium 32, 219–224.
Selvan J. A., Delahoy A. E., Guo S. and Li Y.-L. (2006), ‘A new light-trapping TCO for
nc-Si:H solar cells’, Solar Energy Mater. Solar Cells 90, 3371–3376.
Shah A. (2010), ‘Thin-film Silicon solar cells’, EPFL Press, ISBN 1420066749.
Shah A., Schade H., Vanecek M., Meier J., Vallat-Sauvain E., Wyrsch N., Kroll U., Droz C.
and Bailat J. (2004), ‘Thin-film silicon solar cell technology, Progr. Photovoltaics 12,
113–142.
Sheng P., Bloch A. N. and Stepleman R. S. (1983), ‘Wavelength-selective absorption
enhancement in thin-film solar cells’, Appl. Phys. Lett. 43, 579–581.
Shockley W. and Queisser H. J. (1961), ‘Detailed balance limit of efficiency of p-n junction
solar cells’, J. Appl. Phys. 32, 510–519.
Sichanugrist P., Yoshida T., Ichikawa Y. and Sakai H. (1993), ‘Amorphous silicon oxide
with microcrystalline phase’, J. Non-Cryst. Solids 164–166, 1081–1084.
Smets A. H. M., Kessels W. M. M. and van de Sanden M. C. M. (2007b), ‘The effect of ion-
surface and ion-bulk interactions during hydrogenated amorphous silicon deposition’,
J. Appl. Phys. 102, 073523-1–073523-12.
Smets A. H. M., Matsui T. and Kondo M. (2008), ‘High-rate deposition of microcrystalline
silicon p-i-n solar cells in the high pressure depletion regime’, J. Appl. Phys. 104,
034508-1–034508-11
Smets A. H. M., and van de Sanden M. C. M. (2007a), ‘Relation of the Si-H stretching
frequency to the nanostructural Si-H bulk environment’, Phys. Rev. B 76, 073202-1–
073202-4.
Smith Z. and Wagner S. (1985), ‘Intrinsic dangling-bond density in hydrogenated amorphous
silicon’, Phys. Rev. B 32, 5510–5513.
Söderström K. (2013), ’Coupling light into thin silicon layers for high-efficiency solar cells’,
PhD Thesis, EPFL No. 5714.
Söderström K., Bugnon G., Biron R., Pahud C., Meillaud F., Haug F.-J. and Ballif C. (2012b),
’Thin-film silicon triple-junction solar cell with 12.5% stable efficiency on innovative
flat light-scattering substrate’, J. Appl. Phys. 112, 114503–114503-4.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch04 page 204
Söderström K., Bugnon G., Haug F.-J. and Ballif C. (2012a), ‘Electrically flat/optically
rough substrates for efficiency above 10% in n-i- p thin-film silicon solar cells’, Proc.
MRS Spring Meeting, San Francisco, CA, pp. 39–44.
Söderström K., Escarré J., Cubero O., Haug F.-J., Perregaux S. and Ballif C. (2010), ‘UV
nano imprint lithography technique for the replication of back reflectors for n-i- p
thin-film silicon solar cells’, Progr. Photovoltaics 19, 202–210.
Söderström T., Haug F. J., Niquille X., Terrazoni-Daudrix V. and Ballif C. (2009), ‘Asym-
metric intermediate reflector for tandem micromorph thin-film silicon solar cells’, Appl.
Phys. Lett. 94, 063501–063501–3.
Soppe W. J., Muffler H. J., Biebericher H. C., Devilee C., Burgers A. R., Poruba A., Hodakova
L. and Vanecek M. (2005), ‘Optical and structural properties of microcrystalline silicon
grown by microwave PECVD’, Proc. 20th EU-PVSEC, Barcelona, pp. 1604–1607,
3DV.3.21.
Spear W. and Le Comber P. (1976), ‘Electronic properties of substitutionally doped amor-
phous Si and Ge’, Phil. Mag. 33, 935–949.
Sriraman S., Agarwal S., Aydil E. S. and Maroudas D. (2002), ‘Mechanism of hydrogen
induced crystallization of amorphous silicon’, Nature 418, 62–65.
Staebler D. L. and Wronski C. R. (1977), ‘Reversible conductivity changes in discharge-
produced amorphous Si’, Appl. Phys. Lett. 31, 292–294.
Staebler D. L. and Wronski C. R. (1981), ‘Stability of n-i- p amorphous silicon solar cells’,
Appl. Phys. Lett. 39, 733–735.
Steimen I. (2011), Private communication based on Oerlikon Solar process and LCA anal-
yses performed with EMPA CH.
Stradins P. (2003), ‘Light-induced degradation in a-Si: H and its relation to defect creation’,
Solar Energy Mater. Solar Cells 78, 349–367.
Strahm B. and Hollenstein C. (2010), ‘Powder formation in SiH4 -H2 discharge in large-area
capacitively coupled reactors: a study of the combined effect of interelectrode distance
and pressure’, J. Appl. Phys. 107, 023302-1–023301-7.
Strahm B., Howling A., Sansonnens L. and Hollenstein C. (2007), ‘Plasma silane concen-
tration as a determining factor for the transition from amorphous to microcrystalline
silicon in SiH4 /H2 discharges’, Plasma Sources Sci. Technol. 16, 80–89.
Street R. and Winer K. (1989), ‘Defect equilibria in undoped a-Si: H’, Phys. Rev. B 40,
6236–6249.
Strobel C., Zimmermann T., Albert M., Bartha J. W. and Kuske J. (2009), ‘Productivity
potential of an in-line deposition system for amorphous and microcrystalline silicon
solar cells’, Solar Energy Mater. Solar Cells 93, 1598–1607.
Stutzmann M., Jackson W. B. and Tsai C. C. (1985), ‘Light-induced metastable defects in
hydrogenated amorphous silicon: a systematic study’, Phys. Rev. B 32, 23–47.
Tabuchi T., Takashiri M. and Mizukami H. (2003), ‘Hollow electrode enhanced RF glow
plasma for the fast deposition of microcrystalline silicon’, Surface Coatings Technol.
173, 243–248.
Takagi T., Ueda M., Ito N., Watabe Y. and Kondo M. (2006), ‘Microcrystalline silicon
solar cells fabricated using array-antenna-type very high frequency plasma-enhanced
chemical vapor deposition system’, Japan. J. Appl. Phys., 45, 4003–4005.
Takano A. and Kamoshita T. (2004), ‘Light-weight and large-area solar cell production
technology’, Japan. J. Appl. Phys. 34, 7976–7983.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch04 page 205
Takatsuka H., Takeuchi Y., Yamauchi Y., Shioya T. and Kawai Y. (2005), ‘Production of
large area VHF plasma using ladder-shaped electrode’, Surface Coatings Technol.
2005, 972–975.
Takeuchi Y., Nawata Y., Ogawa K., Serizawa A., Yamauchi Y. and Murata M. (2001),
‘Preparation of large uniform amorphous silicon films by VHF-PECVD using a ladder-
shaped antenna’, Thin Solid Films 386, 133–136.
Taneda N., Oyama T. and Sato K. (2007), ‘Light scattering effects of highly textured trans-
parent conductive oxide films’, Techical Digest 17th PVSEC, Fukuoka, pp. 309–312,
5O-B6-01.
Tawada Y., Okamoto H. and Hamakawa Y. (1981), ‘a-SiC:H/a-Si:H heterojunction
solar cell having more than 7.1% conversion efficiency’, Appl. Phys. Lett. 39,
237–239.
TEL (2014), TEL Solar press release, 9 July 2014.
Terakawa A., Hishida M., Aya Y., Shinohara W., Kitahara A., Yoneda H., Iseki M. and
Tanaka M. (2011), ‘R&D on thin-film silicon solar cells in Sanyo’, Proc. 26th PVSEC,
Hamburg, pp. 2362–2365, 3BO.4.2.
Terasa R., Albert M., Grüger H., Haiduk A. and Kottwitz A. (2000), ‘Investigation of growth
mechanisms of microcrystalline silicon in the very high frequency range’, J. Non-cryst.
Solids 266–269, 95–99.
Tiedje T., Cebulka J., Morel D. and Abeles B. (1981), ‘Evidence for exponential band tails
in amorphous silicon hydride’, Phys. Rev. Lett. 46, 1425–1428.
Torres P., Meier J., Flückiger R., Kroll U., Selvan J., Keppner H., Shah A., Littelwood S.,
Kelly I. and Giannoules P. (1996), ‘Device grade microcrystalline silicon owing to
reduced oxygen contamination’, Appl. Phys. Lett. 69, 1373–1375.
Ulbrich C., Zahren C., Noll J., Lambertz A., Gerber A. and Rau U. (2011), ‘Analysis of the
short-circuit current gains by anti-reflective texture coating on silicon thin-film solar
cells’, Proc. 26th PVSEC, Hamburg, pp. 302–305, 1CV.3.29.
Usui S. and Kikuchi M. (1979), ‘Properties of heavily doped GD-Si with low resistivity’,
J. Non-Cryst. Solids 34, 1–11.
van den Donker M. N., Gordijn A., Stiebig H., Finger F., Rech B., Stannowski B., Bartlb R.,
Halmers E. A. G., Schlatmann R. and Jongerden G. J. (2007), ‘Flexible amorphous and
microcrystalline silicon tandem solar modules in the temporary superstrate concept’,
Solar Energy Mater. Solar Cells 91, 572–580.
van Gestel D., Chahal M., van der Wilt P. C., Qiu Y., Gordon I., Im J. S. and Portmans J.
(2010), ‘Thin-film polycrystalline silicon solar cells with low intragrain defect density
made via laser crystallization and epitaxial growth’, Proc. 35th IEEE PVSC, Honolulu,
HI, pp. 279–282.
Vanecek M., Neykova N., Babchenko O., Purkrt A., Poruba A., Remes Z., Holovsky J.,
Hruska K., Meier J. and Kroll U. (2010), ‘New 3-dimensional nanostructured thin-film
silicon solar cells’, Proc. 5th World PVSEC, Valencia, pp. 2763–2766, 3CO.12.2.
Vepřek S. and Marečeka V. (1968), ‘The preparation of thin layers of Ge and Si by chemical
hydrogen plasma transport’, Solid-State Electronics 11, 683–684.
Vetterl O., Finger F., Carius R., Hapke P., Houben L., Kluth O., Lambertz A., Mück A.,
Rech B. and Wagner H. (2000), ‘Intrinsic microcrystalline silicon: a new material for
photovoltaics’, Solar Energy Mater. Solar Cells 62, 97–108.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch04 page 206
Vogler B., Kerschbaumer J., Kuhn H., Mark A., Poppeller M., Schneider S., Zimin D. and
Zindel A. (2008), ‘TCO 1200◦ C Oerlikon production tool for transparent conductive
oxide thin-films’, Proc. 23rd EU-PVSEC, Valencia, pp. 2492–2493.
Wadia C., Alivisatos A. P. and Kammen D. M. (2009), ‘Materials availability expands the
opportunity for large-dcale photovoltaics deployment’, Environmental Sci. Technol.
43, 2072–2077.
Wang C. and Lucovsky G. (1990), ‘Intrinsic microcrystalline silicon deposited by remote
PE-CVD: a new thin-film photovoltaic material’, Proc. 21st IEEE PVSC, Orlando, FL,
pp. 1614–1618.
Winer K. (1990), ‘Defect formation in a-Si:H’, Phys. Rev. B 41, 12150–12161.
Wittmaack K. (2003), ‘Analytical description of the sputtering yields of silicon bombarded
with normally incident ions’, Phys. Rev. B 68, 235211–235221.
Wronski C. (1996), ‘Amorphous silicon technology: coming of age’, Solar Energy Mater.
Solar Cells 41, 427–439.
Xu X., Zhang J., Beglau D., Su T., Ehlert S., Li Y., Pietka G., Worrel C., Lord K., Yue G.,
Yan B., Banerjee A., Yang J. and Guha S. (2010), ‘High efficiency large-area a-SiGe:H
and nc-Si:H based multi-junction solar cells: a comparative study’, Proc. 5th World
PVSEC, Valencia, pp. 2783–2787, 3CO.13.1.
Yablonovitch E., Allara D., Chang C., Gmitter T. and Bright T. (1986), ‘Unusually low
surface-recombination velocity on silicon and germanium surfaces’, Phys. Rev. Lett.
57, 249–252.
Yablonovitch E. and Cody G. D. (1982), ‘Intensity enhancement in textured optical sheets
for solar cells’, IEEE Trans. Electron Devices, 29, 300–305.
Yamamoto K., Nakajima A., Yoshimi M., Sawada T., Fukuda S., Suezaki T., Ichikawa M.,
Goto M., Meguro T., Matsuda T., Sasaki T. and Tawada Y. (2006), ‘High efficiency
thin-film silicon hybrid cell and module with newly developed innovative interlayer’,
Proc. 4th World PVSEC, Honolulu, HI, pp. 1489–1492.
Yamamoto K., Nakajima A., Yoshimi M., Sawada T., Fukuda S., Suezaki T., Ichikawa M.,
Koi Y., Goto M., Meguro T., Matsuda T., Kondo M., Sasaki T. and Tawada Y. (2004),
‘A high efficiency thin-film silicon solar cell and module’, Solar Energy 77, 939–949.
Yamamoto K., Toshimi M., Suzuki T., Tawada Y., Okamoto T. and Nakajima A. (1998),
‘Thin-film poly-Si solar cell on glass substrate fabricated at low temperature’, Proc.
MRS Spring Meeting, San Francisco, CA, pp. 131–138.
Yamamoto K., Yoshimi M., Tawada Y., Fukuda S., Sawada T., Meguro T., Takata H., Suezaki
T., Koi Y. and Hayashi K. (2002), ‘Large area thin-film Si module’, Solar Energy Mater.
Solar Cells 74, 449–455.
Yan B., Yang J. and Guha S. (2003), ‘Effect of hydrogen dilution on the open-circuit voltage
of hydrogenated amorphous silicon solar cells’, Appl. Phys. Lett. 83, 782–784.
Yan B., Yue G. and Guha S. (2007), ‘Status of nc-Si:H solar cells at United Solar and
roadmap for manufacturing a-Si:H and nc-Si:H based solar panels’, Proc. MRS Spring
Meeting, San Francisco, CA, A15–01.
Yan B., Yue G., Sivec L., Yang J., Guha S. and Jiang C.-S. (2011), ‘Innovative dual function
nc-SiOx :H layer leading to a >16% efficient multi-junction thin-film silicon solar cell’,
Appl. Phys. Lett. 99, 113512-1–113512-3.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch04 page 207
Yang J., Banerjee A., Glatfelter T., Hoffman K., Xu X. and Guha S. (1994), ‘Progress in
triple-junction amorphous silicon based alloy solar cells and modules using hydrogen
dilution’, Proc. 1st World PVSEC, Honolulu, HI, pp. 380–385.
Yang J., Banerjee A. and Guha S. (1997), ‘Triple-junction amorphous silicon alloy solar
cell with 14.6% initial and 13.0% stable conversion efficiencies’, Appl. Phys. Lett. 70,
2975–2977.
Yang L. and Chen L. (1993), “‘Fast” and “slow” metastable defects in hydrogenated amor-
phous silicon’, Appl. Phys. Lett. 63, 400–402.
Yu Z., Raman A. and Fan S. (2010), ‘Fundamental limit of light trapping in grating struc-
tures’, Optics Express 18, A366–A380.
Yue G., Sivec L., Yan B., Yang J. and Guha S. (2010), ‘High efficiency hydrogenated
nanocrystalline silicon based solar cells deposited by optimized Ag/ZnO back reflec-
tors’, Proc. 25th EU-PVSEC, Valencia, pp. 3196–3200, 3AV.2.17.
Yunaz I. A., Hashizume K., Miyajima S., Yamada A. and Konagai M. (2009), ‘Fabrication
of amorphous silicon carbide films using VHF-PECVD for triple-junction thin-film
solar cell applications’, Solar Energy Mater. Solar Cells 93, 1056–1061.
Zeman M., van Swaaij R. A. C. M. M., Metselaar J. W. and Schropp R. E. I. (2000), ‘Optical
modeling of a-Si:H solar cells with rough interfaces: effect of back contact and interface
roughness’, J. Appl. Phys. 88, 6436–6443.
Zhu H., Bunte E., Hüpkes J., Siekmann H. and Huang S. M. (2009), ‘Aluminium doped zinc
oxide sputtered from rotatable dual magnetrons for thin-film silicon solar cells’, Thin
Solid Films 517, 3161–3166.
Zhu H., Kalkan A., Hou J. and Fonash S. (1998), ‘Applications of AMPS-1D for solar cell
simulation’, Proc. AIP Conf. 462, pp. 309–314.
Zimmermann T., Strobel C., Albert M., Beyer W., Gordijn A., Flikweert A. J., Kuske J. and
Bartha J. W. (2010), ‘Inline deposition of microcrystalline silicon solar cells using a
linear plasma source’, Physica Stat. Solidi C 7, 1097–1100.
May 2, 2013 14:6 BC: 8831 - Probability and Statistical Theory PST˙ws
CHAPTER 5
POLYCRYSTALLINE CADMIUM
TELLURIDE PHOTOVOLTAIC DEVICES
TIMOTHY A. GESSERT
National Renewable Energy Laboratory
Golden, Colorado, USA 8040
[email protected]
and
DIETER BONNET
Private Consultant, formerly Antec Solar
5.1 Introduction
Thin-film photovoltaic (PV) devices based on CdTe absorbers represent one of
the fastest growing segments of all PV technologies (Mehata, 2010; Coggeshall
and Margolis, 2011). It is even more remarkable that most of this impact has
occurred within the past few years. Much of the reason for this rapid development
can be traced to two factors: (i) Thin-film PV has been designed specifically to
embody production advantages over historic PV products (i.e. wafer-based tech-
nologies); (ii) CdTe PV presently enjoys production advantages over other thin-film
technologies.
In this chapter, we review the development and present state of CdTe PV
technology to describe the advantages that have made this technology one of two
dominant forces in the present PV market. Although much is understood regarding
the physics and material science of the CdTe thin-film device, we will see that
producing PV modules that are cost-effective in the present market(s) remains a
complex technological and financial undertaking. Furthermore, although current
CdTe commercial technology demonstrates the lowest specific production cost per
watt of any PV technology, we will also suggest that additional improvements and
further market penetration remain likely.
Finally, we will argue that as the economic viability of CdTe PV products
improves further, the impacts of less-obvious considerations will emerge. These
209
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch05 page 210
This benchmark lifetime was an order of magnitude longer than that observed
for the lowest-doped, monocrystalline CdTe. Because the homojunction emitter
would have to be heavily doped to reduce series resistance effects, most research
efforts during the last 40 years have been directed toward development of CdTe
heterojunction devices. Recently, however, reports have emerged suggesting that
monocrystalline homojunction CdTe-based devices may demonstrate a significant
opportunity for multijunction concentrator PV devices (Carmody et al., 2010).
In the CdTe heterojunction configuration, the choice for an appropriate het-
eroface material is limited by several considerations. For example, because of the
near-optimum bandgap of CdTe, the heterojunction should be designed so that
most of the absorption occurs within the CdTe bulk. Thus, the heteroface partner
must act as a highly transparent, low-resistance window layer and not be respon-
sible for carrier generation. Because all known wide-gap, low-resistivity window
materials are n-type, the CdTe layer must be p-type. The choice of window layer
is further constrained in that it is better for it to have a small lattice mismatch with
CdTe to avoid excessive interface recombination. Finally, to promote long-term
stability, the window material should be composed of elements that are slow to
diffuse into CdTe, preferably with the cation from Group II (e.g. Zn, Cd, Hg) to
optimise conduction-band alignment.
Although these considerations provide a guide for the choice of the wide-
bandgap heteroface window, a wide range of materials have been investigated,
including CdS (Bonnet and Rabenhorst, 1972; Yamaguchi et al., 1977; Bube et al.,
1977), ZnO (Aranovich et al., 1977) and ZnSe (Bube et al., 1975). It is notewor-
thy that the 5–6%-efficient CdS/CdTe devices fabricated by Bonnet in 1972 were
not only the first CdS/CdTe devices reported but were also the first all-thin-film
CdS/CdTe devices to be made, in that gas-phase transport was used to deposit thin
films of both layers.
Some of the other window materials may eventually demonstrate advantages
over CdS. However, the most efficient cells produced to date have the CdS/CdTe
configuration (Bube, 1988; Chu, 1988; Wu, 2004; Green, 2013). Additionally,
because the CdS/CdTe configuration represents a 3-element system (rather than
4-element, or more), it may produce fewer potential native defects, and ulti-
mately present fewer unforeseen industrial (production) problems arising from
interdiffusion.
Table 5.1 Historical and present commercial companies involved in CdTe PV.
Approximate CdTe absorber
time period Name Location deposition
1975–1984 Photon Power Inc. El Paso, USA Chemical spraying
1984–1992 Photon Energy Inc. Golden, USA
1992–1997 Golden Photon Inc.
1977–c.1997 Matsushita Corp. Japan Screen-printed sublimation
1980–1984 Monosolar Inc. Santa Monica, USA Electrodeposition
1984–1999 BP Solar London, UK
1999–2002 BP/Solarex Baltimore, USA
1980–1989 AMETEK Harleysville, USA Electrodeposition
1994–2002 ANTEC GmbH Arnstadt, Germany Vacuum sublimation
2002–present Antec Solar Energy
1990–1999 Solar Cells Inc. Toledo, USA Vacuum sublimation
1999–present First Solar
2003–2007 Solar Fields Calyxo Toledo, USA Non-vacuum sublimation
2007–present USA
2006 Ziax Corp. Arvada, USA Vacuum sublimation
2006–2008 Primestar Solar
2008–present Primestar GE Solar
2006–present Solexant Corp. San José, USA Non-vacuum printing
2007–2009 AVA Solar Fort Collins, USA Vacuum sublimation
2009–2012 Abound Solar
2008–present Xunlight 26 Solar Toledo, USA Sputtering
2008–present SunPrint/Alion Inc Richmond, USA Non-vacuum printing
2008–present Advanced Solar Hangzhou, China Vacuum sublimation
Power
2007–present WK Solar Group Toledo, USA Vacuum sublimation
2008–c.2011 Clean Cell Int. Sunnyvale, USA Vacuum sublimation
2009–present REEL Solar Menlo Park, USA Electrodeposition
2010–present Encore Solar Fremont, USA Electrodeposition
In some cases, several entries are listed, indicating name transitions and/or involvements by different
corporate entities. This table has been assembled from various sources, including reports, websites and
personal communications, and is intended to be representative rather than exhaustive.
devices of the early 1970s. Although the only companies presently in commercial
(or pre-commercial) production are using vacuum sublimation, several compa-
nies are now re-exploring the potential advantages of non-vacuum CdTe deposi-
tion processes. Braun and Skinner (2007) provide a more detailed discussion of
the successes and difficulties that can be encountered during CdTe PV module
commercialisation.
Figure 5.1 Schematic diagram showing main functional components of a typical thin-film
CdTe PV solar cell device configured as a superstrate structure. The arrow on the right
indicates the direction and approximate location of the junction electric field.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch05 page 214
and passes through one or more transparent conducting oxide (TCO) layers and a
CdS window layer. It is then absorbed in the CdTe absorber layer.
Because the glass superstrate provides the mechanical surface onto which all
subsequent layers are deposited, the sequence of layer deposition is TCO, then CdS,
then CdTe, and then back contact. Although the superstrate design is currently the
most widely used, and the present world record thin-film CdTe device is a super-
strate structure (Rose et al., 1999; Wu, 2004; Green et al., 2013; Gloeckler et al.,
2013), the substrate design continues to attract both research and commercial inter-
est. Here the device is typically grown on a non-transparent substrate that is often
either a Mo or Mo-coated stainless steel foil. This results in a deposition sequence
opposite to the superstrate device (i.e. back contact, then CdTe, then CdS, then
TCO, then encapsulation). Although the currently reported maximum performance
of substrate devices is significantly lower than that of high-performance superstrate
devices (∼13% efficiency vs. ∼20% efficiency: Romeo et al., 1992; Singh et al.,
1999; Dhereet al., 2012; Duenow et al., 2012; Gretener et al., 2012; Kranz et al.,
2013; Green et al., 2013), the substrate configuration has advantages, in particular in
avoiding the use of glass, which could be exploited in certain commercial products.
These include applications requiring high specific power (i.e. watts/kg), the ability
to produce flexible products for aerospace and/or building-integrated PV markets,
and/or envisioned use of roll-to-roll processing during module manufacture.
Figure 5.2 shows a scanning electron microscopy (SEM) cross-section of a
superstrate CdTe device produced at NREL. This particular device was produced
on barium silicate glass (Corning 7059), and had an SnO2 :F TCO layer and SnO2
buffer layer produced by MOCVD, a CdS layer produced by chemical bath deposi-
tion (CBD), a CdTe layer produced by close-space sublimation (CSS), a dry CdCl2
activation process, a chemical pre-contact etch step and a graphite-paste contact.
Figure 5.2 SEM micrograph showing cross-section of CdS/CdTe superstrate device pro-
duced at NREL. The layers have been colour-enhanced for clarity.
sufficiently high (>324◦C for Cd and >449◦C for Te), liquid elemental regions
can reside on the surface of the crystal. The CdTe itself will melt, as shown by the
outside of the overall curve, and the melting point is a function of off-stoichiometry.
The highest melting point at ∼1092◦ C is for stoichiometric CdTe.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch05 page 216
Figure 5.4 Temperature vs. composition diagram for CdTe showing extent of electrically
active centres for both Cd- and Te- rich concentration. Source: Zanio (1978).
Although single-phase CdTe exists only very near the 50/50% composition, the
material can sustain a Cd or Te deficiency to a small extent. As shown in Fig. 5.4,
this off-stoichiometry can extend to a maximum value of about 5 × 1017 cm−3
before the Te or Cd deficiency will spontaneously produce separate Cd or Te
phases. Recalling that most materials have a density of atoms on the order of
1×1022 cm−3 , this means that the maximum off-stoichiometry that can be sustained
in crystalline CdTe at thermodynamic equilibrium is only about 0.01%. Analysis
of off-stoichiometry also suggests that, for equilibrium conditions, temperatures
greater than 600◦C can yield higher excess Te or Cd than temperatures below 500◦C
(Greenberg, 1996). Depending on the concentration of other defects, the excess Te
or Cd can manifest in polycrystalline films as a material with net acceptors or net
donors. Therefore, if a high degree of stoichiometry control can be maintained
at high growth temperatures, it may be possible to have these defects assist with
various types of junction functionality. However, we shall see that the influence of
the CdCl2 treatment and Cu incorporation during contacting also has a significant
impact on electrical properties.
In addition to a relatively simple diagram space, another important reason
CdTe possesses production advantage is that it demonstrates a property known as
congruent sublimation and condensation. When a source of bulk CdTe is heated to a
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch05 page 217
temperature at which the CdTe surface sublimes into Cd and Te2 gas, the Cd and Te
atoms evaporate at roughly the same rate (i.e. congruent sublimation). Although
the vapour pressure of Cd is higher than that of Te2 at any given temperature,
Te2 has two atoms, so the arrival rates of Te and Cd atoms at a surface are very
similar. Furthermore, when Cd and Te2 vapours condense onto a (cooler) surface,
they again do so at about the same rate (i.e. congruent condensation). These two
congruent processes enable a bulk source of stoichiometric CdTe to evaporate and
condense into a film of nearly stoichiometric CdTe. If for some reason Cd and Te
arrive at a growing film surface at different rates, then because the vapour pressure
of Cd (or Te2 ) above a pure Cd (or Te) surface is much higher than that of either
above a CdTe surface, any depositing Cd or Te surface atoms that do not quickly
coordinate into the CdTe material will re-evaporate from the surface, limiting the
formation of Cd or Te secondary phases.
The previous discussion pertains to high-purity, crystalline CdTe formed
under equilibrium conditions. However, for thin-film PV devices, the CdTe will be
formed using less pure materials and ambients, and relatively non-equilibrium con-
ditions. These differences can result in polycrystalline thin films that can contain
numerous structural and impurity defects that can affect the material properties of
the film. Luckily, because the kinetics of film growth for CdTe are fairly rapid, even
high-rate film growth processes tend to retain much of the structural regularity that
is observed for bulk crystalline materials.
Although CdTe is ambipolar (i.e. it can be doped both n-type and p-type),
the material tends to self-compensate. This means if one attempts to incorporate
acceptors by substituting Group I dopants on the cation [Cd] site, or Group V
dopants on the anion [Te] site, a dopant concentration will be reached where the
lattice will spontaneously create compensating donors, with the result that the
further change in net acceptor concentration is minimal (Zunger, 2003).
Production of epitaxial CdTe films using molecular beam epitaxy (MBE) tech-
niques can achieve n- and p-type doping into the mid-1018 cm−3 range using
extrinsic dopants (Taterenko et al., 1993; Dhese et al., 1994). For polycrystalline
PV devices, the CdTe is typically deposited without extrinsic dopants added. The
as-deposited CdTe is often believed to be p-type as a result of cadmium vacancies
(VCd ); however, the electrical resistance of as-deposited CdTe is very high, so the
electrical properties of as-deposited polycrystalline thin-films are not well estab-
lished. Most device technologists know only that a working junction results if the
CdTe is deposited onto an n-type heteroface layer, such as CdS, and a back contact
is fabricated onto the CdTe. The conclusion is that the CdTe must have been, or
have become, p-type. As we will see in the following discussion, for some CdTe
deposition processes, it is possible that the as-deposited CdTe may be p-type,
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch05 page 218
Temperature stability
CdTe technologists suggest that higher substrate temperatures minimise the defect
density in the CdTe film. Although firm evidence of this has not been reported,
it is true that the devices with higher performance tend to have had the CdTe
deposited at higher temperature. In the case of commonly used soda-lime float glass
substrates, it is essential to keep the processing temperatures below 550–575◦ C to
avoid mechanical distortion due to the low strain and softening points, and to
avoid altering the degree of temper of the as-fabricated glass. For some industrial
processes, the temperature may be much lower to avoid even small distortions in
the glass and/or to limit loss of glass temper. For higher processing temperatures,
borosilicate (and historically, barium silicate/Corning 7059) glass substrates have
been used for laboratory devices to allow studies of material functionality at higher
temperature. Because of the size of the potential market for thin-film CdTe PV,
many commercial glass companies are considering the use of other types of glass
compositions that may demonstrate higher strain points than soda-lime glass, or
may embody other commercial advantages.
cooling. This gives rise to structural defects and can also affect the adhesion of the
film(s). Although this is another area where studies have not been conducted and/or
reported, it remains prudent to choose a glass with a CTE that closely matches that
of the material(s) being deposited.
complete explanation for this difference has not been reported, it partly arises
because the maximum electron mobility of present commercial-grade SnO2 :F is
lower than in other TCOs (in the range of 20–25 cm2 V−1 sec−1 ). This means that
to achieve both low sheet resistance and high optical transmission, higher carrier
concentration must be used in the film (∼7 × 1020 cm−3 ). This higher carrier
concentration will produce more free-carrier absorption, limiting the transmitted
near infrared (NIR) light entering the PV device, and thus the short-circuit current.
Many research groups are presently investigating avenues to improve the per-
formance of SnO2 :F, and/or develop alternative TCOs for CdTe devices (Dhere
et al., 2010). For example, a recent high-performance thin-film CdTe device uses a
Cd2 SnO4 TCO layer (Wu et al., 1996; Coutts et al., 2000; Wu, 2004). An alternative
avenue to improve the transmission of TCOs for PV applications involves forming
TCO alloys with higher dielectric permittivity (Gessert et al., 2011). Because the
dielectric permittivity affects the plasma wavelength, this technique can yield a
TCO with much less NIR absorption.
• The BL allows thinner CdS layers to be used. One proposed reason for this is
that the space charge on the n-type side of the junction can expand into the
BL when the CdS becomes too thin to balance the charge needed to maintain
the space-charge width in the p-type side of the junction. This enables the
space charge in the p-CdTe region to remain sufficiently wide as the CdS
layer is thinned, thereby avoiding device functionality dominated by voltage-
dependent collection (i.e. low fill factor arising from loss of minority-carrier
collection at high forward bias). A different explanation is that the BL reduces
surface recombination at the CdS/TCO interface. The suggested benefit is that
reduced recombination at this interface may improve device performance if
thinner CdS layers position this interface closer to the CdS/CdTe junction
region.
• The BL provides tolerance to shunt (or short) paths from the back contact to
the front contact. These pathways will be present if the CdTe layer contains
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch05 page 221
cracks or pinholes that may form during CdTe deposition, or during pre-contact
treatments such as chemical etching. These shunt paths are particularly prob-
lematic if vacuum-deposited metals are used in the contact (as opposed to
mechanically applied paste contacts).
• Certain types of BL can allow higher temperatures or longer times to be used
during the CdCl2 treatment. The issue here is that during the CdCl2 process
(see Section 5.6.7), delamination can occur between the CdS and TCO layers
if the CdCl2 process temperature is excessively high. Indeed, the optimum
CdCl2 processing conditions are typically defined as the maximum time and
temperature that can be tolerated before delamination. It has been suggested
this benefit may be more likely for a chemically active BL, such as Zn2 SnO4
or related alloys that can form mixed alloys with the CdS.
• A chemically active buffer layer, such as Zn2 SnO4 or related alloys, may
consume some of the CdS layer, reducing the amount of (optically absorbing)
CdS, and producing in its stead a wider bandgap CdZnS material (Wu, 2004).
The particular choices of glass superstrate, TCO, buffer layer and CdS will
significantly affect the amount of light that can enter the junction region of a CdTe
thin-film PV device, and thus will impact the device short-circuit current. This
is illustrated in Fig. 5.5, which compares the quantum efficiency consistent with
the 16.7% CdTe device produced at NREL (i sc =∼ 26 mA cm−2, dotted quantum
efficiency curve) with a historic commercial device (i sc =∼ 19 mA cm−2 , solid
curve) (Wu et al., 2002; Green et al., 2012).
The comparison reveals significant loss differences in both the NIR region
(600–850 nm) and the UV region (300–600 nm). About one half of the ∼3 mA cm−2
NIR loss is believed to be due to absorption in the (high-Fe) commercial glass and
the other half is due to the free-carrier absorption of commercial TCO. For the
NREL device shown, barium silicate glass (Corning 7059) and a Cd2 SnO4 TCO
layer were used to minimise both these NIR losses. It should also be noted that the
commercial TCO modelled for this figure (using Drude Theory approximations)
might actually suggest less NIR absorption than typical commercial SnO2 :F. The
difference in UV loss between the commercial and NREL devices shown in this
figure (∼4 mA cm−2) is primarily due to the use of relatively thick layers (typ-
ically ∼150–300 nm) of CdS in the commercial device. In the NREL device, a
Zn2 SnO4 -alloy BL was used to allow a CdS layer that is typically thinned to ∼70 nm
thick. From this discussion, it is clear that additional performance benefits remain
possible in commercial modules by optimisation of the CdS thickness. However,
this will require improved understanding of the function(s) and optimisation of
specific BLs.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch05 page 222
Figure 5.5 Comparison of quantum efficiency curves from the 16.7% efficiency CdS/CdTe
device at NREL (i sc =∼ 26 mA cm−2 ) and a representative commercial CdS/CdTe device
(i sc =∼ 19 mA cm−2 ). The red triangles indicate the approximate amount of loss in each
region. The figure also shows several different types of glass superstrate and a modelled
absorption curve for a typical commercial TCO.
process. In the case of sputter deposition of CdS, oxygen addition can also lead
to the formation of a nanocrystalline CdS phase (Wu et al., 2002). It is believed
that quantum-confinement effects yield a wider bandgap for nanocrystalline CdS,
allowing increased device short-circuit current from the CdS:O/CdTe device.
Figure 5.6 (a) Cross-section schematic of research-scale CSS system; (b) cross-section of
research-scale gas-phase transport system.
that the maximum grain size tends to be equal to the film thickness. However, the
initially small grains have significant implications for superstrate devices because
the grain size nearest to the junction region (where light absorption occurs) is
generally small, and therefore issues arising from interface and/or grain-boundary
recombination will be more pronounced.
Reported electrical properties of as-deposited polycrystalline CdTe films are
based primarily on resistivity and hot-probe studies (Zanio, 1978; McCandless,
2003; Dhere and Li, 2012). These suggest that low deposition temperatures produce
n-type films with resistivity at or above ∼108 ohm cm. Intermediate-temperature
deposition yields as-deposited resistivity in the range of 107 –108 ohm cm, and
the films tend to be intrinsic but can demonstrate either n- or p-type character-
istics. Research devices that have been produced at high temperatures suggest
as-deposited films are slightly p-type (Dhere et al., 2008). The transition of con-
ductivity type from n to p with increasing temperature may be linked to the forma-
tion of more cadmium vacancy defects. The CdCl2 treatment reduces the resistivity
by 3–4 orders of magnitude and can assist conversion of intrinsic or n-type films
to lightly p-type (this is discussed more fully in Section 5.6.7). Although more
detailed information on the carrier concentration and type of as-deposited films
would be valuable, results of CV and Hall analyses of as-deposited and CdCl2 -
treated CdTe films are generally inconclusive because the low carrier concentration
produces surface and interface depletion regions that are equal to or wider than
typical film thicknesses. Therefore, nearly all of the electrical data reported for
polycrystalline CdTe are for films that have been treated after deposition, typically
by CdCl2 and Cu contacting.
As we note below, the assumed interplay between different process steps is one
of the things that make CdTe a difficult material to process reproducibly. However,
we will also note that improved understanding of any of these process steps may
not only lead to their elimination, but also elimination of related process issues.
Figure 5.7 SEM cross section of CdS/CdTe device illustrating the effect of oxygen in
the CdS on CdS consumption during CdTe deposition: (a) CSS-deposited CdS containing
no oxygen. Note significant consumption of CdS; (b) CSS-deposited CdS that has been
post-deposition treated to incorporate oxygen. Note much less consumption of CdS; (c)
CBD-deposited CdS. Note very little consumption of CdS. Source: Albin et al. (2002).
a CdSTe-alloy region, its function and preferred attributes remain areas of active
debate.
In addition to the CdTe deposition temperature and CdCl2 treatment (McCan-
dless et al., 1999), another important parameter affecting the CdSTe-alloy region
is the amount of oxygen available within this region (Albin et al., 2002). Figure 5.7
shows that the amount of oxygen in the CdS layer alters the interdiffusion between
the CdS and CdTe layers during high-temperature CdTe deposition. In general,
adding oxygen to the CdS layer reduces the amount of CdS consumed during
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch05 page 227
high-temperature CdTe deposition. The reduced CdS consumption not only limits
the extent of the CdSTe alloy that forms on the CdTe side of the device but also
appears to limit the amount of Te that diffuses into the CdS layer.
It has been suggested that the resulting moderation of the extent of the CdSTe
thickness by oxygen limits the depth of the CdSTe/CdTe interface, which limits the
depth of the quasi-homojunction (Dhere et al., 1996; Dhere et al., 2008). One pos-
sible explanation for these effects of oxygen is that oxygen acts as an isoelectronic
substitutional defect for Te and/or S vacancies. The formation of these defects could
limit diffusion, and possibly reduce interface recombination. However, at this time,
the effects of this diffusion on junction formation and ultimate cell performance
are not well understood.
(a) (b)
Figure 5.8 SEM micrographs of CdCl2 oxychloride residuals from wet CdCl2 processing
treatment: (a) low magnification; (b) higher magnification. Source: Gessert et al. (2001).
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch05 page 229
(a)
(b)
Figure 5.9 SIMS analysis of the effect of CdCl2 treatment on S diffusion as a function of
CdTe deposition temperature. (a) 550◦ C; (b) 600◦ C. Source: Dhere et al. (1996).
deposited at lower temperature (550◦ C). Recalling that the amount of oxygen avail-
able at the CdS/CdTe interface also affects the extent of the CdSTe layer, the result
suggests controlling the formation of the CdSTe layer requires (at least) controlling
the CdTe temperature, the CdCl2 treatment temperature and time, and the amount
of available oxygen. The interrelationship between these parameters begins to
explain why developing a reproducible and cost-effective CdTe PV module pro-
cess requires a deep understanding of the mechanisms at work and the ability to
control these mechanisms.
The above description of how the CdCl2 treatment affects CdSTe and CdTe
material properties provides clues to how the treatment also affects device perfor-
mance. Figures 5.10a and 5.10b show the effect of the CdCl2 treatment process
on device performance as a function of CdTe CSS substrate temperature (Dhere
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch05 page 230
Figure 5.10 Effect of a wet CdCl2 treatment on (a) device Voc and (b) i sc as a function of
CSS CdTe substrate temperature. Source: Dhere et al. (1996).
et al., 1996). The figure shows that while higher CdTe deposition temperatures
lead to improved device performance, the CdCl2 process improves performance at
all substrate temperatures. In this data set, a device process parameter (substrate
temperature) was altered to yield a device that responds better to a given (wet)
CdCl2 treatment. However, depending on process constraints, the CdCl2 treatment
could also be varied to produce a higher performance for a given device process.
Because these process alterations can be subtle, and are often made without full
knowledge of what aspect of cell functionality is being altered, it may be difficult to
know precisely how a specific CdCl2 process parameter may be affecting junction
performance.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch05 page 231
been used for the active diffusing dopant species, other Group I species (Au or Ag)
or Group V species (P, As, Sb or Bi) have also demonstrated potential (Zanio, 1978).
While the precise defect formation that occurs during dopant diffusion remains
debated, it is generally accepted that a successful diffusion alters the electrical
properties of the underlying CdTe layer so that it becomes sufficiently p-type to
establish a strong field (i.e. narrow space-charge region) in the ∼0.5–1 µm of the
device nearest the CdS layer (Gessert et al., 2006). Furthermore, Cu diffusion
at appropriate temperatures (∼250–350◦C) has been found to increase electron
lifetime near the CdS/CdTe interface (Gessert et al., 2009; Gessert et al., 2010).
As mentioned in Section 5.6.5, it is possible that the CdCl2 -treated CdTe
layers may be intrinsic or n-type. This suggestion is supported by contact studies
where dopant diffusion can be limited to very low concentrations. These show,
for CdTe deposited at intermediate temperatures (450–575◦C), devices formed
without Cu dopant diffusion from the contact typically demonstrate very little (if
any) rectification (Gessert et al., 2002). However, when even a small amount of
Cu is allowed to diffuse from the contact, strong junction rectification is observed,
suggesting that contact diffusion may be converting the layer from n-type to p-type.
Additional differences in the response of the CdTe layer to contact diffusion
are observed between layers grown at intermediate (450–575◦C) and high temper-
ature (600–625◦C). Capacitance–voltage analysis of devices grown at intermediate
temperatures, shown in Fig. 5.11, indicates that initial dopant diffusion from the
contact causes the junction space charge to contract systematically. This suggests
that even the initial dopant diffusion from the contact leads to an increase in net
acceptor concentration (Gessert et al., 2005; Gessert et al., 2006]. In contrast, con-
tact diffusion into material grown at high temperatures (600–625◦C) causes the
space charge to first expand, and then contract (Fig. 5.11). This suggests that the
net acceptor density decreases with initial dopant diffusion, and thereafter increases
(Chin et al., 2010).
It has recently been reported that this seemingly confusing functionality can
be explained by noting that higher deposition temperature will possibly yield CdTe
layers in which the effective ionisation energy changes with the extent of Cu dif-
fusion (Ma et al., 2011). Specifically, high-temperature deposition yields CdTe
with more cadmium vacancy defects (VCd ) than would be present at intermedi-
ate temperatures. Further, the VCd exists as a double acceptor, with both a lower
(∼130 meV) and a higher ionisation energy (∼210 meV). As Cu diffuses into the
CdTe, it preferentially substitutes onto available VCd sites and becomes an acceptor
CuCd . However, the ionisation energy of CuCd (∼220 meV) is greater than either
of the VCd activation energies, and so the observed net acceptor density (NA )
decreases. This reduction in NA continues until the available cadmium vacancies
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch05 page 233
are consumed. At this point, remaining Cu can either displace the Cd in the CdTe
lattice (the formation energy of CuCd in CdTe is ∼1.31 eV) to form additional CuCd
acceptors, or form Cu interstitial donors (Cui ) (Wei and Zhang, 2002). Recent stud-
ies have shown, depending on chemical potential (i.e. whether CdTe is formed with
excess Cd or Te concentration), the CuCd defect formation is preferred over Cui
up to a certain defect concentration, at which point formation of the compensating
Cui will commence and pin the Fermi level (Ma et al., 2011). Further, depending
on the particular thermal process conditions following Cu diffusion, the Cui can
begin to dominate, type-converting the material back to n-type.
In contrast to doping the CdTe layer from the contact and producing the n − p
junction functionality described above, an alternative contact design is intended to
yield a low-resistance CdTe/CIFL interface while sustaining a CdTe layer that is
electrically intrinsic (i.e. not extrinsically doped p-type) while still demonstrating
a long minority-carrier lifetime (i.e. n −i − p junction functionality) (Sites and Pan,
2007). For this junction design, if the intrinsic CdTe layer demonstrates sufficiently
long lifetime, it may also be advantageous for the CIFL to function as an electron
reflector. Modelling suggests wider-bandgap tellurides or their alloys (e.g. ZnTe,
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch05 page 234
(a)
(b)
Figure 5.12 (a) Diagram illustrating the effect of band bowing and phase separation in
the CdS x Te1−x layer. The circle at x =∼ 0.1 and 620◦ C indicates the most likely sulphur
concentration, based on the effective bandgap and the available single-phase region; (b)
diagram illustrating the location of the junction in a CdS/CdTe PV device from a historic
(top) and more recent (bottom) perspective.
the stoichiometry of the CdSTe alloy, this can exist as a single-phase, hexagonal
material (wurtzite, ∼85–100% sulphur), a single-phase cubic material (zincblende,
0–∼15% sulphur), or a mixed-phase CdSTe material (between ∼15 and ∼85%
sulphur) (Nunoue et al., 1990). During intermixing of the CdS and CdTe layers
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch05 page 236
at the CdTe deposition temperature, it is believed that all three of these phases
will be present in what has so far been described simply as the CdSTe layer.
Many technologists believe that it is reasonable to assume the CdSTe layer will be
primarily in the wurtzite (i.e. S-rich) phase nearest the CdS layer, primarily in the
zincblende (Te-rich) phase nearest the CdTe layer, and in the mixed phase between
these two (primarily) single-phase endpoints. However, only a few studies of this
region have been reported (Dhere et al., 2008).
For the type of devices produced at NREL, we believe the electrical junction
forms between the non-alloyed CdTe, and a diffusion-formed zincblende CdSTe
(Dhere et al., 2008). Analysis of the IR absorption characteristics of quantum
efficiency data indicates the effective sulphur concentration in the junction region is
∼10%. This is estimated by calculating the effective bandgap from device quantum
efficiency (∼1.46 eV), while accounting for the known band bowing (Compaan
et al., 1996). This estimation process is shown graphically in Fig. 5.12a. Two
schematic diagrams illustrating both the historic and more recent understanding of
the layer structure, as well as the associated electric fields, are shown in Fig. 5.12b.
5.8.1 Reliability
As any PV technology matures from the laboratory to commercial production, a
point is reached when product reliability must be understood well enough to assign a
product warranty. Presently, most CdTe thin-film PV module manufacturers like to
align module warranties with established PV alternatives (particularly wafer-based
crystalline Si). This level of reliability is typically consistent with the panel deliv-
ering at least 80% of its initial power after 25 years of field deployment. Assuming
a constant exponential degradation, this equates to about 0.9% power loss per
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch05 page 237
5.9 Conclusions
This chapter has presented a technological description of components and process
alternatives consistent with the present generation of CdTe thin-film PV devices. We
have also suggested why this technology may embody considerable opportunity for
becoming a significant part of future large-scale electricity production. However,
even when economic requirements are met, the establishment of a sustainable
energy-production infrastructure from CdTe technologies will require additional
consideration of longer-term issues. These include reliability, materials availability,
environmental considerations and energy payback time.
Acknowledgements
The authors wish to thank R. G. Dhere, D. S. Albin, T. M. Barnes, J. N. Duenow,
S.-H. Wei and T. J. Coutts of NREL for assistance and thoughtful suggestions
regarding the preparation of this chapter. Portions of this work were supported
under DOE Contract No. DE-AC36-08-GO28308 to NREL.
References
Albin D. S., Demtsu S. H. and McMahon T. J. (2006), ‘Film thickness and chemical pro-
cessing effects on the stability of cadmium telluride solar cells’, Thin Solid Films 515,
2659–2668.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch05 page 240
Albin D. S., Yan Y. and Al-Jassim M. M. (2002), ‘The effect of oxygen on interface
microstructure evolution in CdS/CdTe solar cells’, Progr. Photovoltaics 10, 309–322.
Aranovich J., Golmayo D., Fahrenbruch A. L. and Bube R. H. (1980), ‘Photovoltaic prop-
erties of ZnO–CdTe heterojunctions prepared by spray pyrolysis’, J. Appl. Phys. 51,
4260–4268.
Artobolevskaya E. S., Afanaseva E. A., Vodopyanov L. K. and Sushkov V. P. (1967), ‘Pho-
toelectromagnetic effect in cadmium telluride’, Sov. Phys. Semicond. 1, 1531.
Basol B. M. (1984), ‘High-efficiency electroplated heterojunction solar cell’, J. Appl. Phys.
55, 601–603.
Bell R. O., Serreze H. B. and Wald F.V. (1975), ‘A new look at CdTe solar cells’, Proc. 11th
IEEE Photovoltaic Specialists Conf., Scotsdale, AZ, pp. 497–502.
Bonnet D. and Rabenhorst H. (1972), ‘New results on the development of thin film p-
CdTe/n-CdS heterojunction solar cell’, Proc. 9th IEEE Photovoltaic Specialists Conf.,
Silver Springs, MD, pp. 129–132.
Braun G. W. and Skinner D. E. (2007), ‘Experience scaling-up manufacturing of emerging
photovoltaic technologies’, Natl. Renewable Energy Laboratory Subcontract Report,
NREL/SR-640-39165, January 2007.
Bube R. H. (1988), ‘CdTe junction phenomena’, Solar Cells 23, 1–17.
Bube R. H., Buch F., Fahrenbruch A. L., Ma Y. Y. and Mitchell K. W. (1977), ‘Photovoltaic
energy conversion with n-CdS- p-CdTe heterojunctions and other II–VI junctions’,
IEEE Trans. Electron. Dev. ED-24, 487–492.
Bube R. H., Fahrenbruch A., Aranovich J., Buch F., Chu M. and Mitchell K. (1975), ‘Applied
research in II–VI compound materials for heterojunction solar cells’, NSF Report No.
NSF/RANN/SE/AER-75-1679/75/4.
Carmody M., Mallick S., Margetis J., Kodama R., Biegala T., Xu D., Bechmann P., Garland
J. W. and Sivananthan S. (2010), ‘Single-crystal II–VI on Si single-junction and tandem
solar cells’, Appl. Phys. Lett. 96, 153502–153502-3.
Chin K. K., Gessert T. A. and Wei S.-H. (2010), ‘The roles of Cu impurity states in CdTe
thin film solar cells’, Proc. 35th IEEE Photovoltaic Specialists Conf., Piscataway, NJ,
pp. 1915–1918.
Chu T. L. (1988), ‘Cadmium telluride solar cells’, in Coutts T. J. and Meakin J. D., (eds),
Vol. 3 of Current Topics in Photovoltaics, Academic Press, New York.
Chu T. L., Chu S. S., Britt J., Ferekides C., Wang C., Wu C. Q. and Ullal H. S. (1992),
‘14.6% Efficient thin-film cadmium telluride heterojunction solar cell’, IEEE Electron
Dev. Lett. 13, 303–304.
Coggeshall C. and Margolis R. (2011), Solar Vision Study, U.S. Dept. of Energy Report.
Compaan D., Feng Z., Contreas-Puente G., Narayanswamy C. and Fisher A. (1996),
‘Properties of pulsed laser deposited CdSx Te1−x films on glass’, MRS Proc. 426,
367–372.
Corwine C., Sites J. R., Gessert T. A., Metzger W. K., Dippo P., Li J., Duda A. and Teeter
G. (2005), ‘CdTe photoluminescence: comparison of solar-cell material with surface-
modified single crystals’, Appl. Phys Lett. 86, 221909–221909-3.
Coutts T. J., Young D. L. and Li X. (2000), ‘Characterization of transparent conducting
oxides’, MRS Bulletin 25, 58–65.
Cusano D. A. (1963), ‘CdTe solar cells and photovoltaic heterojunctions in II–IV Com-
pounds’, Solid-State Electron. 6, 217–218.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch05 page 241
Cusano D. A. (1967), ‘Thin film studies and electro-optical effects’, in Aven M. and Prener
J. S. (eds.), Physics and Chemistry of II–VI Compounds, North-Holland Publishing
Co., Amsterdam, Ch. 14, pp. 709–766.
Cusano D. A. and Lorenz M. R. (1964), ‘CdTe hole lifetime from the photovoltaic effect’,
Solid State Commun. 2, 125–128.
Dhere R. G., Albin D. S., Rose D. H., Asher S. E., Jones K. M., Al-Jassim M. M., Moutinho
H. R. and Sheldon P. (1996), ‘Intermixing at the CdS/CdTe interface and its effect on
device performance’, MRS Proc. 426, 361–366.
Dhere R. G., Bonnet-Eymard M., Charlet E., Peter E., Duenow J. N., Li J. V., Kuciauskas D.
and Gessert T. A. (2011), ‘CdTe solar cell with industrial Al:ZnO on soda-lime glass’,
Thin Solid Films 519, 7142–7145.
Dhere R. G., Duenow J. N., DeHart C. M., Li J. V., Kuciauskas D. and Gessert T. A.,
(2012), ‘Development of substrate structure CdTe photovoltaic devices with perfor-
mance exceeding 10%’, Proc. 38th IEEE Photovoltaic Specialists Conf., IEEE, Piscat-
away, NJ, pp. 3208–3211.
Dhere R. G. and Li X. (2012), NREL, private communication.
Dhere R. G., Zhang Y., Romero M. J., Asher S. E., Young M., To B., Noufi R. and Gessert
T. A. (2008), ‘Investigation of junction properties of CdS/CdTe solar cells and their
correlation to device properties’, Proc. 33rd IEEE Photovoltaic Specialists Conf., San
Diego, CA, Manuscript No. 279.
Dhese K. A., Devine P., Ashenford D. E., Nicholls J. E., Scott C. G., Sands D. and Lunn B.
(1994), ‘Photoluminescence and p-type conductivity in CdTe:N grown by molecular
beam epitaxy’, J. Appl. Phys. 76, 5423–5428.
Duenow J. N., Dhere R. G., Kuciauskas D., Li J. V., Pankow J. W., Dippo P. C., DeHart C.
M. and Gessert T. A., ‘Oxygen incorporation during fabrication of substrate CdTe pho-
tovoltaic devices’, Proc. 38th IEEE Photovoltaic Specialists Conf., IEEE, Piscataway,
NJ, pp. 3225–3229.
Fthenakis V. M. (2004), ‘Life cycle impact analysis of cadmium in CdTe PV production’,
Renewable and Sustainable Energy Reviews 8, 303–334.
Fulop G., Doty M., Meyers P., Betz J. and Liu C. H. (1982), ‘High-efficiency electrodeposited
cadmium telluride solar cells’, Appl. Phys. Lett. 40, 327–328.
Garcia D. A., Entine G. and Tow D. E. (1974), ‘Detection of small bone abscesses with a
high-resolution cadmium telluride probe’, J. Nucl. Med. 15, 892–895.
Gessert T. A., Asher S., Johnston S., Duda A., Young M. R. and Moriarty T. (2006), ‘For-
mation of ZnTe:Cu/Ti contacts at high temperature for CdS/CdTe devices’, Proc. 4th
World Conf. Photovoltaic Energy Conversion, pp. 432–435.
Gessert T. A., Asher S., Johnston S., Young M., Dippo P. and Corwine C. (2007), ‘Analysis
of CdS/CdTe devices incorporating a ZnTe:Cu/Ti contact’, Thin Solid Films 515, 6103–
6106.
Gessert T. A., Burst J., Li X., Scott M. and Coutts T. J. (2011), ‘Advantages of transparent
conducting oxide thin films with controlled permittivity for thin film photovoltaic solar
cells’, Thin Solid Films 519, 7146–7148.
Gessert T. A., Dhere R. G., Duenow J. N., Kuciauskas D., Danevce A. and Bergeson J. D.
(2010), ‘Comparison of minority carrier lifetime measurements in superstrate and sub-
strate CdTe PV devices’, Proc. 37th IEEE Photovoltaic Specialists Conf., IEEE, Pis-
cataway, NJ, pp. 335–339.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch05 page 242
Gessert T. A., Metzger W. K., Dippo P., Asher S. E., Dhere R. G. and Young M. R. (2009),
‘Dependence of carrier lifetime on Cu-contacting temperature and ZnTe:Cu thickness
in CdS/CdTe thin film solar cells’, Thin Solid Films 517, 2370–2371.
Gessert T. A., Perkins C. L., Asher S. E., Duda A. and Young M. R. (2003), ‘Study of
ZnTe:Cu/metal interfaces in CdS/CdTe photovoltaic solar cells’, Mater. Res. Soc.
Symp., Vol. 796, pp. 79–84.
Gessert T. A., Romero M. J., Johnston S., Keys B. and Dippo P. (2002), ‘Spectroscopic
cathodoluminescence studies of the ZnTe:Cu contact process for CdS/CdTe solar cells’,
Proc. 29th IEEE Photovoltaic Specialists Conf., IEEE, Piscataway, NJ, pp. 535–538.
Gessert T. A., Romero M. J., Perkins C. L., Asher S. E., Matson R., Moutinho H. and
Rose D. (2001), ‘Microscopic residuals on polycrystalline CdTe following wet CdCl2
treatment’, MRS Proc. 668, H1.10.1–H1.10.6.
Gessert T. A., Smith S., Moriarty T., Young M., Asher S., Johnston S., Duda A. and DeHart C.
(2005), ‘Evolution of CdS/CdTe device performance during Cu diffusion’, Proc. 31st
IEEE Photovoltaic Specialists Conf., IEEE, Piscataway, NJ, pp. 291–294.
Gloeckler M., Sankin I. and Zhao Z. (2013), ‘CdTe solar cells at the threshold to 20%
efficiency’, IEEE J. Photovoltaics 3, 1389–1398.
Green M. A., Emery K., Hishikawa Y., Warta W. and Dunlop E. D. (2012), ‘Solar cell
efficiency tables (Version 39)’, Progr. Photovoltaics 20, 12–20.
Green M. A., Emery K., Hishikawa Y., Warta W. and Dunlop E. D. (2013), ‘Solar cell
efficiency tables (Version 42)’, Progr. Photovoltaics 21, 827–837.
Greenberg J. H. (1996), ‘P–T–X phase equilibrium and vapor pressure scanning of non-
stoichiometry in CdTe’, J. Cryst. Growth 161, 1–11.
Gretener C., Perrenoud J., Kranz L., Kneer L., Schmitt R., Buecheler S. and Tiwari N. (2012),
‘CdTe/CdS thin film solar cells grown in substrate configuration’, Progr. Photovoltaics
21, 1580–1586.
Gur I., Fromer N. A., Geier M. L. and Alivisatos A. P. (2005), ‘Air-stable all-inorganic
nanocrystal solar cells processed from solution’, Science 310, 462–465.
Halliday D. P., Potter M. D. G., Boyle D. S. and Durose K. (2001), ‘Photoluminescence
characterization of ion implanted CdTe’, MRS Proc.668, H1.8.1.
Held M. and Ilg R. (2008), ‘Life cycle assessment (LCA) of CdTe thin film PV modules and
material flow analysis (MFA) of cadmium within EU27’, Proc. 23rd Eur. Photovoltaic
Solar Energy Conf., Sept. 1–5, Valencia.
Hsiao K. J. and Sites J. R. (2009), ‘Electron reflector strategy for CdTe solar cells’, Proc.
34th IEEE Photovoltaic Specialists Conf., Fort Collins, CO, pp. 1846–1850.
Jordan J. F. (1993), ‘Photovoltaic cell and method’, U.S. Patent No. 5261968.
Karpov V. G., Shvydka D., Roussillon Y. and Compaan A. D. (2003), ‘The mesoscale physics
of large-area photovoltaics’, Proc. 3rd World Conf. Photovoltaic Energy Conversion,
WCPEC–3, Osaka, pp. 495–498.
Kester J. L., Albright S., Kaydanov V., Ribelin R., Woods L. M. and Phillips J. A. (1996),
‘CdTe solar cells: electronic and morphological properties’, AIP Conf. Proc. Ser. 394,
162–169.
Kranz L., Schmitt R., Gretener C., Perrenoud J., Pianezzi F., Uhl A. R., Keller D., Buecheler
S. and Tiwari A.N. (2013), ‘Progress towards 14% efficient CdTe solar cells in substrate
configuration’, Proc. 39th IEEE Photovoltaic Specialists Conf., Tampa, FL, 1644–
1648.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch05 page 243
Levi D., Albin D. and King D. (2000), ‘Influence of surface composition on back-contact
performance in CdTe/CdS PV devices’, Progr. Photovoltaics 8, 591–602.
Li X., Gessert T. A., Matson R. J., Hall J. F. and Coutts T. J. (1993), ‘Microstructural study
of sputter-deposited CdTe thin films’, J. Vac. Sci. Technol. A 12, 1608–1613.
Li X., Niles D. W., Hasoon F. S., Matson R. J. and Sheldon P. (1999), ‘Effect of nitric–
phosphoric acid etches on materials properties and back-contact formation in CdTe-
based solar cells’, J. Vac. Sci. Technol. A 17, 805–809.
Ma J., Wei S.-H., Gessert T. A. and Chin K. K. (2011), ‘Carrier density and compensation
in semiconductors with multiple dopants and multiple transition energy levels: case of
Cu impurities in CdTe’, Phys. Rev. B 83, 245207–245214.
McCandless B. (2003), ‘Effects of treatments on CdTe film conductivity’, NREL Subcon-
tract Report #ADJ-1-30630-12, 23 December.
McCandless B. E., Youm I. and Birkmire R. W. (1999), ‘Optimization of vapor post-
deposition processing for evaporated CdS/CdTe solar cells’, Progr. Photovoltaics 7,
21–30.
Mehata S. (2010, ‘PV news annual data collection result: 2010 cell, module production
explodes past 20 GW’, PV News 29, May 2010.
Meyer E., Martini M. and Sternberg J. (1972), ‘Measurement of the disappearance rate of
Se-75 sodium selenite in the eye of the rat by a CdTe medical probe’, IEEE Trans.
Nucl. Sci. 19, 237–243.
Moutinho H. R., Dhere R. G., Al-Jassim M. M., Levi D. H. and Kazmerski L. L. (1999),
‘Investigation of induced recrystallization and stress in close-spaced sublimated and
radio-frequency magnetron sputtered CdTe thin films’, J. Vac. Sci. Technol. A 17,
1793–1798.
Naumov G. P. and Nikolaeva O. V. (1961), ‘The efficiency of transformation of direct solar
radiation energy into electric energy using a CdTe photocell’, Soviet Phys. Solid State
3, 2718.
Nicoll F. H. (1963), ‘The use of close spacing in chemical-transport systems for growing
epitaxial layers of semiconductors’, J. Electrochem. Soc. 110, 1165–1167.
Nunoue S.-Y., Hemmi T. and Kato E. (1990), ‘Mass spectrometric study of the phase bound-
aries of the CdS–CdTe system’, J. Electrochem. Soc., 137, 1248–1251.
Powell R. C., Dorer G. L., Reiter N. A., McMaster H. A., Cox S. M. and Kahle T. D.
(1999), ‘Apparatus and method for depositing a material on a substrate’, U.S. Patent
No. 5945163.
Raugei M., Bargigli S. and Ulgiati S. (2007), ‘Life cycle assessment and energy pay-back
time of advanced photovoltaic modules: CdTe and CIS compared to poly-Si’, Energy
32, 1310–1318.
Romeo N., Bosio A. and Canevari V. (1992), ‘Large crystalline grain CdTe thin films for
photovoltaic application’, Int. J. Solar Energy 12, 183–186.
Rose D. H (1997), ‘The effects of oxygen on CdTe–absorber solar cells deposited by close-
spaced sublimation’, PhD Thesis, University of Colorado, Boulder, CO.
Rose D. H., Hasoon F. S., Dhere R. G., Albin D. S., Ribelin R. M., Li X. S., Mahathongdy Y.,
Gessert T. A. and Sheldon P. (1999), ‘Fabrication procedures and process sensitivities
for CdS/CdTe solar cells’, Progr. Photovoltaics 7, 331–340.
Sasala R. A., Powell R. C., Dorer G. L. and Reiter N. (1996), ‘Recent progress in CdTe
solar cell research at SCI’, AIP Conf. Proc Ser. 394, 171–186.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch05 page 244
Singh V. P., McClure J. C., Lush G. B., Wang W., Wang X., Thompson G. W. and Clark
E. (1999), ‘Thin film CdTe-CdS heterojunction solar cells on lightweight metal sub-
strates’, Solar Energy Mater. Solar Cells 59, 145–161.
Sites J. and Pan J. (2007), ‘Strategies to increase CdTe solar-cell voltage’, Thin Solid Films
515, 6099–6102.
Suyama N., Arita T., Nishiyama Y., Ueno N., Kitamura S. and Murosono M. (1990),
‘CdS/CdTe solar cells by the screen-printing–sintering technique’, Proc. 21st IEEE
Photovoltaic Specialists Conf., Kissimmee, FL, pp. 498–503.
Tatarenko S., Bassani F., Saminadayar K., Cox R. T., Jouneau P. H. and Megnea N. M.
(1993), ‘Indium doping of (001), (111) and (211) CdTe layers grown by molecular
beam epitaxy’, J. Cryst. Growth 127, 318–322.
Tyan Y.-S. and Perez-Albuerne E. A (1982), ‘Efficient thin-film CdS/CdTe solar cells’, Proc.
16th IEEE Photovoltaic Specialists Conf., San Diego, CA, pp. 794–800.
Tyan Y.-S., Vazan F. and Barge T. S. (1984), ‘Effect of oxygen on thin-film CdS/CdTe solar
cells’, Proc. 17th IEEE Photovoltaic Specialists Conf., Kissimmee, FL, pp. 840–844.
Valdna V., Hiie J. and Gavrilov A. (2001), ‘Defects in Cl-doped CdTe thin films’, Solid State
Phenom. 80, 155–162.
Vodakov Y. A., Lomakina G. A., Naumov G. P. and Maslakovets Y. P. (1960), ‘Properties
of p − n junctions in cadmium telluride photocells’, Soviet Phys. Solid State 2, 11.
Waters D. M., Niles D., Gessert T., Albin D., Rose D. and Sheldon P. (1988), ‘Surface analysis
of CdTe after various pre-contact treatments’, Proc. 2nd World Conf. Photovoltaic Solar
Energy Conversion, Vienna, European Commission, Luxembourg, p. 1031.
Wei S.-H. and Zhang S. B. (2002), ‘Chemical trends of defect formation and doping limit
in II–VI semiconductors’, Phys. Rev. B 66, 155211–155221.
Wendt R., Fischer A., Grecu D. and Compaan A. D. (1998), ‘Improvement of CdTe solar cell
performance with discharge control during film deposition by magnetron sputtering’,
J. Appl. Phys. 84, 2920–2925.
Wu X. (2004), ‘High efficiency polycrystalline CdTe thin-film solar cells’, Solar Energy 77,
803–814.
Wu X., Dhere R. G., Yan Y., Romero M. J., Shang Y., Zhou J., DeHart C., Duda A., Perkins
C. and To B. (2002), ‘High efficiency polycrystalline CdTe thin-film solar cells with an
oxygenated amorphous CdS (a-CdS:O) window layer’, Proc. 29th IEEE Photovoltaic
Specialists Conf., New Orleans, LA, pp. 531–535.
Wu X., Keane J. C., Dhere R. G., DeHart C., Albin D. A., Duda A., Gessert T. A., Asher
A. S., Levi D. H. and Sheldon P. (2002) ‘16.5% Efficient CdS/CdTe polycrystalline
thin-film solar cell,’ Proc. 17th European Photovoltaic Solar Energy Conf., October
22–26, 2001,WIP-Renewable Energies, Munich, pp. 995–1000.
Wu X., Mulligan W. P. and Coutts T. J. (1996), ‘Recent developments in RF sputtered
cadmium stannate films’, Thin Solid Films 286, 274–276.
Yamaguchi K., Nakayama N., Matsumoto H. and Ikegami S. (1977), ‘CdS–CdTe solar cell
prepared by vapor phase epitaxy’, Jpn. J. Appl. Physics 16, 1203–1211.
Zanio K. (1978), Cadmium Telluride, Vol. 13 in Semiconductors and Semimetals, Academic
Press, New York.
Zunger A. (2003), ‘Practical doping principles’, Appl. Phys. Lett. 83, 57–59.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch06 page 245
CHAPTER 6
6.1 Introduction
From the early days of photovoltaics until today, thin-film solar cells have always
competed with technologies based on single-crystal materials such as Si and GaAs.
Owing to their amorphous or polycrystalline nature, thin-film solar cells always
suffered from power conversion efficiencies lower than those of the bulk technolo-
gies. This drawback was and still is counterbalanced by several inherent advantages
of thin-film technologies. As in the early years of photovoltaics, space applications
were the driving force for the development of solar cells, the argument in favour
of thin films was their potential lighter weight as compared to bulk materials.
An extended interest in solar cells as a source of renewable energy emerged
in the mid-1970s as the limitations of fossil energy resources were widely recog-
nised. For terrestrial power applications the cost arguments and the superior energy
balance strongly favoured thin films. However, from the various materials under
consideration since the 1950s and 1960s, only three thin-film technologies, namely
amorphous (a-)Si and the polycrystalline heterojunction systems CdS/CdTe, and
CdS/CuInSe2 , have achieved industrial production in any volume.
CuInSe2 was synthesised for the first time by Hahn in 1953 (Hahn et al., 1953).
In 1974, this material was proposed as a photovoltaic material (Wagner et al., 1974)
245
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch06 page 246
window layer deposition (Section 6.3). This section also discusses options which
can be used to design the electronic properties of the absorber material as well
as basic technologies for module production; (iii) the electronic properties of
the finished heterostructure and some methods of analysing them (Section 6.4);
(iv) finally, Section 6.5 discusses the photovoltaic potential of wide-gap chalcopy-
rites, namely CuGaSe2 and CuInS2 , as well as that of the pentenary alloy system
Cu(In,Ga)(S,Se)2 and the possibility of building graded-gap structures with these
alloys. Regarding In and Ga as rare materials has stimulated the development of
the kesterite compounds, in which indium and gallium are replaced by a pair of
Group II and Group IV elements such as Zn–Sn.
This chapter can only very briefly cover those scientific issues that are rele-
vant for photovoltaic applications. For other important points and for more detailed
information, we refer the reader to the literature. More about the structural prop-
erties of Cu(In,Ga)Se2 can be found, for example, in Shay and Wernick (1975),
Kazmerski and Wagner (1985), Coutts et al. (1986), Rockett and Birkmire (1991),
Schock (1996), Bube (1998) and Rau and Schock (1999). Interface properties
of Cu(In,Ga)Se2 and related compounds have been reviewed by Scheer (1997).
For up-scaling and module technologies see, for example, Dimmler and Schock
(1996), and for economic aspects, see Zweibel (1995). A comprehensive overview
of Cu(In,Ga)Se2 (and CdTe)-based photovoltaics is further given in the book by
Scheer and Schock (2011).
Se
Se In
Zn Cu
(a) (b)
Figure 6.1 Unit cells of chalcogenide compounds. (a) Sphalerite or zinc-blende structure
of ZnSe (two unit cells); (b) chalcopyrite structure of CuInSe2 . The metal sites in the two
unit cells of the sphalerite structure of ZnSe are alternately occupied by Cu and In in the
chalcopyrite structure.
2.4 CuGaS2
2.2
2.0
1.8
Eg /eV CuGaSe2
1.6
CuInS2
1.4
1.2 CuInSe 2
1.0
5.4 5.5 5.6 5.7 5.8
a/Angstroms
CuGaSe2 and Zn0.5 Cd0.5 Se that of CuInSe2 ). This difference is because the Cu 3d
band, together with the Se 4 p band, forms the uppermost valence band in the Cu-
chalcopyrites, which is not so in II–VI compounds. However, the system of copper
chalcopyrites covers a wide range of bandgap energies E g from 1.04 eV in CuInSe2
up to 2.4 eV in CuGaS2 , covering most of the visible spectrum. The kesterite com-
pounds Cu2 ZnSnS4 and Cu2 ZnSnSe4 have similar bandgaps to the corresponding
chalcopyrite compounds CuInS2 and CuInSe2 . Figure 6.2 summarises lattice con-
stants a and bandgap energies E g of this system. Any desired alloys between these
compounds can be produced, as there is no miscibility gap in the entire system. We
will discuss the status and prospects of this system in more detail in Section 6.5.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch06 page 249
CuIn5Se8
Se CuIn3Se5
10
Cu2In4Se 7
90
Cu3In5Se9
80 20
CuInSe2
70 30
CuSe2 In2Se3
60 40
In6Se 7
CuSe 50 50 InSe
Cu3Se2 40 60
Cu2Se 70
20 80
10 90
Cu In
90 80 70 60 50 40 30 20 10
Cu11In9
β+ δ
δ α+δ
600
α+
Temperature T/°C
Cu 2 Se(HT)
β
400
α+ β α
α+
200 Cu2 Se(RT)
15 20 25 30
Cu content (at%)
single-phase region DTA cooling
two-phase region DTA heating
material: (i) the ability to dope Cu(In,Ga)Se2 with native defects; (ii) the structural
tolerance to large off-stoichiometries; and (iii) the electrically neutral nature of the
structural defects. It is obvious that the explanation of these effects significantly
contributes to the explanation of the photovoltaic performance of this material. It
is known that the doping of CuInSe2 is controlled by intrinsic defects. Samples
with p-type conductivity are grown if the material is Cu-poor and annealed under
high Se vapour pressure, whereas Cu-rich material with Se deficiency tends to be
n-type (Migliorato et al., 1975; Noufi et al., 1984). Thus, the Se vacancy VSe is
considered to be the dominant donor in n-type material (and also the compensat-
ing donor in p-type material), and the Cu vacancy VCu the dominant acceptor in
Cu-poor p-type material.
Theoretical considerations
By calculating the metal-related defects in CuInSe2 and CuGaSe2 , Zhang et al.
(1998) found that the defect formation energies for some intrinsic defects are so
low that they can be heavily influenced by the chemical potential of the components
(i.e. by the composition of the material) as well as by the electrochemical potential
of the electrons. For VCu in Cu-poor and stoichiometric material, the calculated
formation energy actually becomes negative. This would imply the spontaneous
formation of large numbers of these defects under equilibrium conditions. Low
(but positive) formation energies are also found for the Cu-on-In anti-site CuIn in
Cu-rich material (this defect is a shallow acceptor which could be responsible for
the p-type conductivity of Cu-rich, non-Se-deficient CuInSe2 ). The dependence
of the defect formation energies on the electron Fermi level could explain the
strong tendency of CuInSe2 to self-compensation and the difficulties of achieving
extrinsic doping. The work of Zhang et al. (1998) provides a good theoretical basis
for the calculation of defect formation energies and defect transition energies,
which exhibit good agreement with experimentally obtained data.
Further important results in Zhang et al. (1997) are the formation energies
of defect complexes such as (2Vcu ,InCu ), (CuIn ,InCu ) and (2Cui ,CuIn ), where Cui
is an interstitial Cu atom. These formation energies are even lower than those of
the corresponding isolated defects. Interestingly, (2Vcu ,InCu ) does not exhibit an
electronic transition within the forbidden gap, in contrast to the isolated InCu -anti-
site, which is a deep recombination centre. As the (2Vcu ,InCu ) complex is most
likely to occur in In-rich material, it can accommodate a large amount of excess In
(or likewise deficient Cu) and, at same time, maintain the electrical performance
of the material. Furthermore, ordered arrays of this complex can be thought as
the building blocks of a series of Cu–In–Se compounds such as CuIn3 Se5 and
CuIn5 Se8 (Zhang et al., 1997).
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch06 page 252
Additionally, the microscopic reasons for metastable changes have been inten-
sively investigated in recent years. The (Vcu ,VSe ) divacancy complex (Lany and
Zunger, 2006) and the InCu anti-site defect (Lany and Zunger, 2008) have been iden-
tified as sources for the metastabilities. Since both defects exist in multiple charge
states their influence on the electronic behaviour of CIGS solar cells is rather com-
plex. For example, even in equilibrium the divacancy complex has three different
charge configurations, namely (Vcu ,VSe )+ , (Vcu ,VSe )− , and (Vcu ,VSe )3− , depen-
dent on the position in the band diagram of the ZnO/CdS/CIGS heterostructure
(Urbaniak and Igalson, 2009; Siebentritt et al., 2010).
Device-relevant defects
Let us now concentrate on the defects experimentally detected in photovoltaic
grade (and thus In-rich) polycrystalline films. In-rich material is in general highly
compensated, with a net acceptor concentration of the order of 1016 cm−3 . The
shallow acceptor level VCu (which lies about 30 meV above the valence band)
is assumed to be the main dopant in this material. As compensating donors,
the Se-vacancy VSe as well as the double donor InCu are considered. The most
prominent defect is an acceptor level about 270–300 meV above the valence band,
which has been reported by several groups from deep-level transient spectroscopy
(Igalson and Schock, 1996) and admittance spectroscopy (Schmitt et al., 1995;
Walter et al., 1996b). This defect is also present in single crystals (Igalson et al.,
1995).
As an example, Fig. 6.5 displays a defect density spectrum obtained from
admittance spectroscopy by the method of Walter et al. (1996a). The transition
at ∼300 meV exhibits a broadened energy distribution with a tail in the defect
density towards larger energies. This tail-like distribution is best described by a
characteristic energy E*, as shown in Fig. 6.5. This defect is detected, not only in
In-rich, but in equal amounts also in Cu-rich polycrystalline materials (Herberholz,
1998). An assignment of this defect to the CuIn anti-site is in agreement with the
theoretical calculations of Zhang et al. (1998) as well as with the proposition of
several experimentalists. The importance of this transition derives from the fact that
its concentration is related to the open-circuit voltage of the device (Herberholz
et al., 1997a) and that the defect seems to be involved in the defect metastability
(Igalson and Schock, 1996) (cf. Section 6.4.5).
The lower-energy transition in Fig. 6.5 is attributed to interfacial defects rather
than to a bulk defect (Herberholz et al., 1998) because its activation energy can vary
between 50 meV and 250 meV depending on air-annealing prior to the measurement
(Rau et al., 1999a). Thus, the activation energy of this transition measures the depth
∆E Fn from the vacuum level of the (electron) Fermi level and the conduction-band
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch06 page 253
18
10
interface defect N 1
E
Defect density Dt /cm–3 eV–1
E
17 bulk defect N 2 E
10 E
16
10
high-energy tail
D t ~ exp (–E/E*)
15
10
0 .1 0.2 0.3 0.4 0.5 0.6
Activation energy E / eV
Figure 6.5 Defect density spectrum obtained from admittance spectroscopy of a ZnO–
CdS–CuInGaSe2 heterojunction. The peaks N1 and N2 can be related to interface and bulk
defects (see inset).
energy at the Cu(In,Ga)Se2 surface (Herberholz et al., 1998), as shown in the top
right-hand of Fig. 6.5.
• The film quality has been substantially improved by the crystallisation mech-
anism induced by the presence of Cu y Se (y < 2). This process is further
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch06 page 254
ZnO:Al
i-ZnO
CdS
Cu(In,Ga)Se2
Mo
glass
The effect of these four items on the electronic properties and performance
of Cu(In,Ga)Se2 solar cells will be considered in detail below, as we discuss the
preparation of a Cu(In,Ga)Se2 solar cell step by step.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch06 page 255
Re-evaporation Vapour
Condensation
CuSe Liquid
Growth
CuInSe2 Solid
CuSe
CuInSe2
Mo
Substrate
Figure 6.7 Schematic illustration of the growth of a Cu(In,Ga)Se2 film under Cu-rich con-
ditions. A quasi-liquid Cu–Se phase acts as a flux in a vapour-liquid solid growth mechanism.
during the growth process of co-evaporated films (Klenk et al., 1993). This model
for the growth of Cu(In,Ga)Se2 in the presence of a quasi-liquid surface film of
Cu y Se is highlighted in Fig. 6.7. For Cu(In,Ga)Se2 prepared by selenisation, the
role of Cu2−y Se is similar (Probst et al., 1996); therefore growth processes for
high-quality absorber materials have to go through a copper-rich stage and end
with an indium-rich composition.
Co-evaporation processes
The absorber material yielding the highest efficiencies is Cu(In,Ga)Se2 with
a Ga/(Ga + In) content of ∼20%, prepared by co-evaporation from elemental
sources. Figure 6.8 sketches a co-evaporation set-up as used for the preparation of
laboratory-scale solar cells and mini-modules. The process requires a maximum
substrate temperature of ∼550◦ C for a certain time during film growth, preferably
towards the end of growth. One advantage of the evaporation route is that material
deposition and film formation are performed during the same processing step. A
feed-back loop based on a quadrupole mass spectrometer or an atomic absorption
spectrometer controls the rate of each source. The composition of the deposited
material with regard to the metals corresponds to their evaporation rates, whereas
Se is always evaporated in excess. This precise control over the deposition rates
allows for a wide range of variations and optimisations with different sub-steps or
stages for film deposition and growth. These sequences are defined by the evapo-
ration rates of the different sources and the substrate temperature during the course
of deposition. Figure 6.9 illustrates some of the possibilities, starting with a simple
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch06 page 257
single-step process where all rates as well as the substrate temperature are kept
constant during the whole process (Fig. 6.9a).
Advanced preparation sequences always include a Cu-rich stage during the
growth process and end up with an In-rich overall composition in order to combine
the large grains of the Cu-rich stage with the otherwise more favourable electronic
properties of the In-rich composition. The first example of this kind of procedure
is the so-called Boeing or bilayer process (Mickelsen and Chen, 1980), which
starts with the deposition of Cu-rich Cu(In,Ga)Se2 and ends with an excess In
rate, as illustrated in Fig. 6.9b. Another possibility is the inverted process, where
first (In,Ga)2 Se3 (likewise In, Ga, and Se from elemental sources to form that
compound) is deposited at a lower temperatures (typically around 300◦ C). Then
Cu and Se are evaporated at an elevated temperature until an overall composition
close to stoichiometry is reached (Kessler et al., 1992). This process leads to
smoother film morphology than the bilayer process. The most successful version
of the inverted process is the so-called three-stage process (Gabor et al., 1994)
shown in Fig. 6.9c This process puts the deposition of In, Ga, and Se at the end of
an inverted process to ensure the overall In-rich composition of the film even if the
material is Cu-rich during the second stage. The three-stage process currently leads
to the best solar cells. Variations of the Ga/In-ratio during deposition, as shown in
Fig. 6.9d, allow the design of graded-bandgap structures (Gabor et al., 1996).
Selenisation processes
The second class of absorber preparation routes is based on the separation of
deposition and compound formation into two different processing steps. High effi-
ciencies are obtained from absorber prepared by the selenisation of metal precursors
in H2 Se (Binsma and Van der Linden, 1982; Chu et al., 1984; Kapur et al., 1987)
and by rapid thermal processing of stacked elemental layers in a Se atmosphere
(Probst et al., 1996). These sequential processes have the advantage that large-
area deposition techniques such as sputtering can be used for the deposition of
the materials. The Cu(In,Ga)Se2 film formation then requires a second step, the
selenisation.
The very first large-area modules were prepared by the selenisation of metal
precursors in the presence of H2 Se some 25 years ago (Mitchell et al., 1988). A
modification of this process provided the first commercially available Cu(In,Ga)Se2
solar cells, manufactured by Siemens Solar Industries. This process is schemati-
cally drawn in Fig. 6.10. First, a stacked layer of Cu, In and Ga is sputter-deposited
on the Mo-coated glass substrate. Then selenisation takes place under H2 Se. To
improve device performance, a second thermal process under H2 S is added, result-
ing in an absorber that is Cu(In,Ga)(S,Se)2 rather than Cu(In,Ga)Se2 .
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch06 page 259
In
Cu
DEPOSITION
Mo
glass
heating
H2 Se
SELENISATION
H2S inlet
Figure 6.10 Illustration of the sequential process. Stacked metal layers are selenised and
converted into CuInSe2 in an H2 Se atmosphere.
A variation of this method that avoids the use of the toxic H2Se during selenisa-
tion is the rapid thermal processing of stacked elemental layers (Probst et al., 1996).
Here the precursor includes a layer of evaporated elemental Se. The stack is then
selenised by a rapid thermal process (RTP) in either an inert or a Se atmosphere.
The highest efficiencies are obtained if the RTP is performed in an S-containing
atmosphere (either pure S or H2 S).
On the laboratory scale, the efficiencies of cells made by these preparation
routes are smaller by about 3% (absolute) as compared with the record values.
However, on the module level, co-evaporated and sequentially prepared absorbers
have about the same efficiency. Sequential processes need two or even three stages
for absorber completion. These additional processing steps may counterbalance
the advantage of easier element deposition by sputtering. Also the detailed and
sophisticated control over composition and growth achieved during co-evaporation
is not possible for the selenisation process. Fortunately, the distribution of the
elements within the film grown during the selenisation process turns out to be
close to what one could think to be an optimum, especially if the process includes
the sulphurisation stage. Since the formation of CuInSe2 is much faster than that of
CuGaSe2 , and because film growth starts from the top, Ga is concentrated towards
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch06 page 260
There are many explanations for the beneficial impact of Na, and it is likely
that the incorporation of Na results in a variety of consequences. During film
growth, the incorporation of Na leads to the formation of NaSex compounds. This
slows down the growth of CuInSe2 and could at the same time facilitate the incor-
poration of Se into the film (Braunger et al., 1998b). Also the widening of the
existence range of the α-(CuInSe2 ) phase in the phase diagram, discussed above,
as well as the reported larger tolerance to the Cu/(In + Ga) ratio of Na-containing
thin films, could be explained in this picture. Furthermore, the higher conductivity
of Na-containing films could result from the diminished number of compensat-
ing Vse donors. Wolf et al. (1998) investigated the influence of Na incorporation
on CuInSe2 film formation from stacked elemental layers by means of thin-film
calorimetry. The addition of Na inhibits the growth of CuInSe2 at temperatures
below 380◦C. The retarded phase formation is responsible for the better morphol-
ogy in the case of Na-containing samples.
Another explanation put forward by Kronik et al. (1998) is that Na pro-
motes oxygenation and passivation of grain boundaries. This could account for
the observed enhancement of the net film doping by Na incorporation, through the
diminished positive charge at the grain boundaries. It has in fact been observed
that the surfaces of Na-containing films are more prone to oxygenation than are
Na-free films (Braunger et al., 1998a).
The above explanations deal with the role of Na during growth. However, the
amount of Na in device-quality Cu(In,Ga)Se2 films is of the order 0.1 at.%, which
is a concentration of 1020 cm−3 (Niles et al., 1997), and one may ask the question:
where are these tremendous quantities of Na in the finished absorber? The elec-
tronic effect, i.e. the change of effective doping resulting from Na incorporation,
is achieved at concentrations of ∼1016 cm−3 , four orders of magnitude below the
absolute Na content. It has long been believed that the main part of the Na is situ-
ated at the film surface and the grain boundaries. Final evidence for this hypothesis
was obtained by Niles et al. (1999) with the help of high spatial resolution Auger
electron spectroscopy.
Heske et al. (1996, 1997) investigated the behaviour of Na on the surface of
polycrystalline Cu(In,Ga)Se2 films by X-ray photoelectron spectroscopy (XPS).
They found two different species of Na: (i) The first, denoted ‘reacted’, was
observed on the air-exposed sample or after storing the sputter-cleaned sample
for three days in an ultra-high vacuum (UHV). The second, denoted ‘metallic’,
was found on clean samples either after annealing at 410 K in UHV or after delib-
erate Na deposition from a metallic source. The latter species is considered to be
the active one during crystal growth. In addition, Heske et al. found an increase of
band bending of ∼150 meV induced by the deposition of Na. This finding, as well
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch06 page 262
Influence of oxygen
Air annealing has been an important process step, crucial for good efficiency,
especially of the early solar cells based on CuInSe2 . Also, though often not men-
tioned explicitly, an oxygenation step is still used for most of the present-day
high-efficiency devices. The beneficial effect of oxygen was explained within the
defect chemical model of Cahen and Noufi (1989). In this model, the surface
defects at grain boundaries are positively charged Se vacancies VSe (Fig. 6.11a)
During air annealing, these sites are passivated by O atoms (Fig. 6.11b). Because
of the decreased charge at the grain boundary, the band bending and the recombina-
tion probability for photogenerated electrons are reduced. The surface donors and
their neutralisation by oxygen are important for the free Cu(In,Ga)Se2 surface as
well as for the formation of the CdS/Cu(In,Ga)Se2 interface (Kronik et al., 2000).
Electrical analysis of oxidised and unoxidised samples revealed the validity of the
Cahen–Noufi model for the earlier CdS/CuInSe2 devices (Sasala and Sites, 1993),
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch06 page 263
Ec
(a) Eg
Ev
(b)
Eg
Figure 6.11 Band diagram of the conduction- and valence-band energies across a sin-
gle grain of Cu(In,Ga)Se2 . (a) The electronic states at the grain boundaries are positively
charged. This surface charge is compensated by the negative charges in the depleted grain.
This induces the band-bending electronic states at the grain boundaries shown at the left-
hand grain boundary, and the defect chemical equivalent, dangling bonds, shown at the
right-hand boundary; (b) oxygen passivates these dangling bonds and reduces the band
bending.
Ec
EF Ev EF EF
∆Ev
∆Ev
Figure 6.13 Models for the CuInSe2 surface: (a) Segregation of a Cu-poor CuIn3 Se5 ODC
on the surface; (b) band bending due to surface charges; (c) band bending induces Cu-
depletion at the surface, creating a surface defect layer. The valence-band energy is lowered
due to Cu depletion. The defect layer provides an internal barrier Φi to the electron transport
(Niemegeers et al., 1998).
defect layer model of Niemegeers et al. (1998) and Herberholz et al. (1999). The
free surfaces of as-grown Cu(In,Ga)Se2 films exhibit two prominent features:
• The valence band-edge energy E v lies above the surface-Fermi level E F by
about 1.1 eV for CuInSe2 films (Schmid et al., 1993). This energy is larger than
the bandgap energy E ga of the bulk of the absorber material. This was taken
as an indication of a widening of bandgap at the surface of the film. For the
surfaces of Cu(In1−x Gax )Se2 thin films it was found that E F − E v = 0.8 eV
(almost independent of the Ga content if x > 0) (Schmid et al., 1996b).
• The surface composition of Cu-poor CuInSe2 , as well Cu(In,Ga)Se2 films,
corresponds to a surface composition of (Ga + In)/(Ga + In + Cu) of about 0.75
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch06 page 265
for a range of bulk compositions of 0.5 < (Ga + In)/(Ga + In + Cu) < 0.75.
Both observations have led to the assumption that a phase segregation of
Cu(In,Ga)3Se5 , the so-called Ordered Defect Compound (ODC), occurs at the
surface of the films. The segregation of this β-phase would be compatible
with the phase diagram (see Fig. 6.3) and, as this material displays n-type
conductivity, could yield the explanation for the surface type inversion.
However, the defect layer model considers the surface layer not as n-type bulk
material (as does the ODC model) but as a p+ -layer (cf. Fig. 6.13c). Furthermore,
the defect layer is viewed not as the origin, but rather as the consequence of the
natural surface type inversion. In contrast to the ODC model, surface states are
responsible for the band bending. In turn, this band bending leads to the liberation
of Cu from its lattice sites and to Cu migration towards the neutral part of the
film (Herberholz et al., 1999). The remaining copper vacancies V− Cu close to the
surface result in a high density of acceptor states, i.e. the p+ -defect layer at the film
surface.
nominally undoped (intrinsic) i -ZnO layer (usually of thickness 50–70 nm) and
then by heavily doped ZnO. This three-step process appears at the moment to be
mandatory for high-efficiency devices, but a convincing explanation of the need
for such a relatively complicated three-layer structure, especially the role of the
i -ZnO, is not available at the moment.
step in a chemical bath. Promising materials to replace CdS are In(OH,S), Zn(OH,S)
and ZnSe. However, all these materials require additional precautions to be taken for
the preparation of the absorber surface or front electrode deposition. Dry methods
like evaporation or sputtering of (In,S) or ZnS are developed as a prerequisite for
inline fabrication of cells without breaking the vacuum.
ZnO
(1 µm)
CdS
(0.05 µm) CIGS (2 µm)
Mo (0.5 µm)
1st patterning
2nd patterning
3rd patterning
Figure 6.14 Interconnect scheme and patterning of a Cu(In,Ga)Se2 (CIGS) based module.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch06 page 269
to ensure that the front ZnO of one cell is connected to the back Mo contact of the
next one. Three different patterning steps are necessary to obtain this connection.
The first interrupts the Mo back contact by a series of periodical scribes and
thus defines the width of the cells, which is about 0.5–1 cm. A laser is normally
used to pattern the Mo in this way. The second patterning step is performed after
absorber and buffer deposition, and the final one after window deposition. Scribing
of the semiconductor layer is performed by mechanical scribing or laser scribing.
The total width of the interconnect depends not only on the scribing tools, but also
on the reproducibility of the scribing lines along the entire module. The typical
interconnect width is of the order of 300 µm. Thus, about 3–5% of the cell area
must be sacrificed to the interconnects.
Module fabrication
Figure 6.15 shows the typical sequence for the production of a Cu(In,Ga)Se2 mod-
ule. The technologies for absorber, buffer and window deposition used for module
production are the same as those discussed above for the production of small labo-
ratory cells. However, the challenge of modules is to transform the laboratory-scale
technologies to much larger areas. Basically, the scheme in Fig. 6.15 applies to both
of the two concepts currently used for large-area absorber preparation, selenisation
Mo sputtering
laser patterning
absorber
deposition
chemical
deposition of
CdS
patterning
ZnO deposition
patterning
Figure 6.15 Process sequence for the fabrication process of a Cu(In,Ga)Se2 module.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch06 page 270
multiple sources
coated
raw plates
plates
Figure 6.16 In-line evaporation system for the deposition of CuInSe2 thin film. The line
source evaporates the material from top to bottom (Dimmler and Schock, 1996).
are about 13%. For reaching competitive cost the production volume has to
exceed 100 MW/year; Solar Frontier has established a capacity of about 1GW/year
(Kushiya, 2013).
Ec
E E
E
E Fn
EF
EV
E
Figure 6.17 Band diagram of the CIGS heterojunction showing the conduction and valence
p
band-edge energies E c and E v . The quantities Φb and Φbn are the barriers for holes and
electrons, ∆E Fn is the energy distance between the Fermi level and the conduction-band
energy at the CdS/CuIn(Ga)Se2 heterointerface, ∆E vab and ∆E cab are the valence-band
and conduction-band discontinuities at the buffer/absorber interface, and w p and wn are the
widths of the space-charge regions in the p-type absorber and the n-type window/buffer
respectively.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch06 page 272
offset between the CdS buffer and the ZnO window layer was determined by Ruckh
et al. (1994b) to be 0.4 eV.
50
Cu(In,Ga)Se2 48 α–1 >1.25 µm
α–1 > 0.5µm
40
46
∆isc=
isc/mA cm–2
–2
30 CuInSe2 3 mA cm
44
20 CuInS 2 42
∆isc=0.9 mAcm –2
1.1 1.2
CuGaSe 2
10
ZnO
CdS
0
1.0 1.5 2.0 2.5 3.0 3.5
Bandgap energy Eg/eV
Figure 6.18 Short-circuit current density from the AM 1.5 solar spectrum and correspon-
dence to the bandgap energies of various chalcopyrite compounds, of a typical Cu(In,Ga)Se2
alloy (E g = 1.12 eV), and of the heterojunction partners. The inset shows the losses that
occur if less than 1.25/0.5 µm material is available for light absorption.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch06 page 274
where Io is the incident irradiance and α the absorption coefficient. For direct
semiconductors, α depends on the photon energy hν according to
hν − E g
α(hν) = a (6.2)
hν
The absorption coefficient of Cu(ln,Ga)Se2 with a low Ga content is reason-
ably described by Eq. 6.2 and a ≈ 8 × 104 eV1/2 cm−1 . By reorganising Eq. 6.2
we can calculate the excess energy
hv 2
∆hv = hv − E g = α (6.3)
a
of photons that have an absorption coefficient larger than a given α. For
instance, with absorption lengths L α = α −1 = 1.255 µm and 0.5 µm, we have
∆hν = 10 meV and 62.5 meV. The inset in Fig. 6.18 shows that the losses ∆i sc cor-
responding to the photons that are not absorbed within 1.25 or 0.5 µm of CuInSe2
are 0.9 mA cm−2 and 3.0 mA cm−2, respectively. The lower-energy photons are
either absorbed at the back-metal/absorber interface or reflected out of the cell.
Thus a typical absorber of thickness 1.5 µm absorbs all the light from the solar
spectrum except for a negligible remnant corresponding to less than 1 mA cm−2 .
Next, we have to recognise that light absorbed in the ZnO window layer does
not contribute to the photocurrent. This loss affects absorption and photogeneration
for photons of energy >3.2 eV (the bandgap energy of ZnO). As shown in Fig. 6.18,
this loss of high-energy photons costs about 1.3 mA cm−2. In addition, photons in
the energy range hν < 1.4 eV are absorbed by free carriers within the conduction
band of the window material. High-efficiency heterojunction solar cells avoid free-
carrier absorption by optimising the conductivity of the front electrode, not by a
high concentration, but by the high mobility of the free carriers. In any case, for
wide-gap absorbers with E g > 1.4 eV this loss can be neglected.
Another portion of the solar light is absorbed in the buffer layer. If, for instance,
a CdS buffer layer caused a sharp cut-off of the spectral response at the bandgap
of 2.4 eV, only a total of 38 mA cm−2 or 35.5 mA cm−2 would be available for
the short-circuit current of a CuInSe2 or Cu(In,Ga)Se2 (E g = 1.11 eV) solar cell,
respectively. However, measurements of the external quantum efficiency (EQE)
of a typical ZnO/CdS/Cu(In,Ga)Se2 heterostructure reveal that the EQE typically
drops by a factor of only ∼0.8 in the wavelength range between the bandgap of
CdS and that of the ZnO window layer. About 70–80% of the photons in the
wavelength range 440–510 nm contribute to i sc because the thin buffer layer does
not absorb all photons and about 50% of the electron-hole pairs created in the
buffer layer still contribute to the photocurrent (the hole recombination probability
at the buffer/absorber interface is relatively low (Engelhardt et al., 1999)).
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch06 page 275
MoSe2
Cu(In,Ga)Se2
CSd
i–ZnO
Mo
A′
C B A qV
φpb
The basic equations for the recombination processes (A–C) can be found in
Bube (1992). All recombination current densities i rec for processes A–C can be
written in the form of a diode law
qV
i rec = i o exp −1 (6.4)
βkT
where V is the applied voltage, β the diode quality factor, and kT/q the thermal
voltage. The saturation current density i o is in general a thermally activated quantity
and may be written in the form
Ẽ a
i o = i oo exp (6.5)
kT
where Ẽ a is the activation energy and the prefactor i oo is only weakly temperature-
dependent. The quantities i o and β depend on the details of each recombination
mechanism. Since mechanisms A–C are connected in parallel, the strongest one
will dominate the recombination loss.
At open circuit, the total recombination current density i rec exactly compen-
sates the short-circuit current density i sc . Hence we can write the open-circuit
voltage as
β Ẽ a βkT i oo Ea βkT i oo
Voc = − ln = − ln (6.6)
q q i sc q q i sc
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch06 page 277
where we have assumed that Voc > 3βkT /q, which allows us to consider only the
exponential term in Eq. (6.4). We have also replaced the activation energy Ẽ a by
E a = β Ẽ a , which will prove in the following to be the ‘true’ activation energy of
the carrier recombination processes.
We now discuss the recombination processes A–C in more detail.
Eqs (6.7) and (6.9) on the doping density NA is equal in that an increase of NA by
one order of magnitude yields an increase of Voc of ∆Voc = (kT ln 10)/q ≈ 60 mV.
However, improving the open-circuit voltage by increasing the doping density is
limited by the increased Auger recombination in the QNR and the enhancement
of tunnelling in the SCR (see Green, 1996a and below).
Note that in Eq. (6.9) the activation energy E a is given by E a = A Ẽ a = 2 Ẽ a =
2(E g /2) = E g whereas in the diode equation for space-charge recombination, the
saturation current density is i o ∝ exp(E g /2kT ) and the activation energy is only
E g /2. This demonstrates that we have to correct the activation energies obtained
from, for example, Arrhenius plots of the saturation current density for the effect of
non-ideal diode behaviour in order to obtain the activation energy relevant to Voc .
In Fig. 6.20 we display the open-circuit voltage limitations given by Eqs (6.7)
and (6.9) for a Cu(In,Ga)Se2 solar cell with an absorber layer thickness of
1.5 µm, a bandgap energy E g of 1.11 eV and a short-circuit current density i sc
of 35.4 mA cm−2. The top and the bottom axes, showing the electron diffusion
length L e and the lifetime τe , are connected by L e = (De τe )1/2 and a diffusion
constant which is here assumed to be De = 2.59 cm2 s−1 . As the open-circuit
voltages in Eqs (6.7) and (6.9) can be shifted by the bandgap energy, we have used
the right-hand axis of Fig. 6.20 to display the difference (E g /q) − Voc .
Diffusion length Le / µm
0.5 1 5
Sb= 10 2 cm s –1 0.40
0.70
Open-circuit voltage Voc
0.45
E g /q – Voc (V)
0.50
0.60
0.55
0.55
0.60
0.1 1 10 100
Lifetime τe /ns
Figure 6.20 Correlation of the open-circuit voltage with lifetimes and diffusion lengths for
a device with a bandgap of 1.12 eV. Solid lines are the results of Eq. (6.9) for L e < 0.7 µm.
For L e > 0.7 µm, Eq. (6.7) holds if effective diffusion lengths that take back-surface recom-
bination into account are introduced. The lines with symbols are the results of a complete
device simulation.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch06 page 279
1 1 2Sg
= + (6.10)
poly
τeff τeb g
where τeb is the minority-carrier lifetime within the grain volume and g denotes the
grain size. With the help of Eq. (6.10), we can still use Eqs (6.8) and (6.10) if we
poly
also use the effective diffusion length L eff for polycrystalline materials, given by
−2
+ 2Sg /(Dn g)]−1/2
poly
L eff = [(L mono
eff ) (6.11)
Interface recombination
Here we consider the simple case of an inverted interface, as shown in Fig. 6.13.
If the recombination centres are not too close to the conduction band, the recom-
bination rate is dominated by the concentration p|i f of free holes at the interface.
From the band diagram and taking the interface recombination velocity for holes
p
as S p and the diffusion potential at the p-side of the heterojunction as VD , the
recombination current density is given by
p
p
qV D + ξ Φb − q V
i rec = qS p p|i f = qS p Nv exp − = qS p Nv exp −
kT kT
(6.14)
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch06 page 281
Tunnelling
In the presence of high electrical fields, interface recombination as well as space-
charge recombination may be enhanced by tunnelling. As shown in Fig. 6.19, the
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch06 page 282
charge carriers do not have to overcome the entire energetic barrier, but only a part
of it, in order to recombine. We denote this transport path as tunnelling-enhanced
recombination because it can be described in very similar terms to the classical
recombination mechanism merely by using modified recombination rates. A useful
quantity within the theory of thermally assisted tunnelling is the tunnelling energy
q h̄
E 00 = (NA,o s /m∗)1/2 (6.17)
2
The tunnelling energy E 00 (Padovani and Stratton, 1966) depends only on the
material parameters NA , the permittivity o s and the effective tunnelling mass m ∗ .
First we consider the case of tunnelling-enhanced bulk recombination in the
SCR. Here the modified recombination rate for tunnelling-assisted recombination
can be integrated over the width of the space-charge region. For an exponential
distribution of trap states, a convenient form of the diode law gives the diode quality
factor as (Rau, 1999)
2
E 00
1 1 T
= 1+ ∗ − (6.18)
β 2 T 3(kT )2
The recombination current density is again expressed by Eq. (6.12), but with the
quality factor given by Eq. (6.18).
The case of tunnelling-enhanced interface recombination can be treated by
analogy with the thermionic field emission theory of Schottky contacts (Padovani
and Stratton, 1966). The recombination current density is
√ p
πq Vb (x)E 00 qV −Φb ξ 1
i rec = qS p Nv exp exp exp −1
kT cosh(E 00 /kT ) βkT βkT kT β
(6.19)
1.2
Eg = 1.24 eV
0.8
0.6
0.4
Figure 6.21 The temperature dependence of the open-circuit voltage under various illumi-
nation intensities extrapolates at 0 K to the bandgap energy.
Temperature T/K
300 200 100
– 20 CuInSe 2
Ea =1.04 eV
ln (io /mA cm –2 )
–40
Cu(In,Ga)Se2
– 60 Ea =1.21 eV
– 80
CuGaSe2
–100 Ea =1.67 eV
0.00 4 0 .0 06 0 .0 08 0.01
Inverse temperature 1/T (1/K)
(a)
0.7 CuInSe 2
E00 = 3.9 meV
Inverse quality factor 1/
0.6
0.5
Cu(In,Ga)Se 2
E 00 = 6.5 meV
0.4
CuGaSe2
0.3 E00 = 16 meV
10 0 150 2 00 25 0 300
Temperature T/K
(b)
Figure 6.22 (a) Arrhenius plots of the reverse saturation current multiplied by the diode
quality factor. Straight lines indicating an activation energy equal to the respective bandgap
energy are evidence for bulk recombination; (b) the experimental data of the inverse quality
factor vs. temperature can be fitted to the model that includes tunnelling.
Figure 6.22a compares such modified Arrhenius plots of the saturation current
densities i o derived from three different Cu(In,Ga)Se2 based heterojunctions. The
three curves stem from absorbers with different Ga contents x = 0, 0.28 and 1.
The activation energies E a = 1.04 eV, 1.21 eV and 1.67 eV extracted from the plots
correspond to the bandgap energies of the respective absorber materials, indicating
that recombination in all these devices occurs in the volume of the absorber.
The corrected Arrhenius plots yield the activation energy of the recombination
process whether or not the transport processes are enhanced by tunnelling. More
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch06 page 285
results from the diode law alone. Thus ηfill depends on temperature as well as on the
0
quality factor of the diode. Effects from series resistance Rs and shunt resistance
Rsh also contribute to the fill factor losses. A good approximation is given by
x
x v oc + 0.7 ηfill
ηfill = ηfill 1 − (6.24)
voc rsh
where ηfill
x
= ηfill
0 (1 − r ), r = R i /V and r = R i /V . The description of
s s s sc oc sh sh sc oc
Cu(In,Ga)Se2 solar cells in terms of Eqs (6.23) and (6.24) works reasonably well.
For example, the typical high-efficiency cell of Contreras et al. (1999) has a fill
factor of 79% and the values calculated from Eqs (6.23) and (6.24) are 78.0% (Voc =
678 mV, i sc = 35.2 mA cm−2, Rs = 0.2 Ω cm2 , Rsh = 104 Ω cm2 , β = 1.5).
The dependence of the fill factor on the quality factor (bearing in mind that
this also determines the open-circuit voltage) highlights the importance of this
parameter for the output power of the solar cell. The diode quality factor βL obtained
from the illuminated current–voltage curve is often different from βD of the dark
curve, with βL > βD . One explanation for this important fact could be the finite
barrier Φbn for the electron transport across the CdS/Cu(In,Ga)Se2 heterointerface.
Looking at Fig. 6.17, a band offset ∆E cab ≈ 0.3 eV between the absorber and
the buffer layer seems to represent a substantial barrier hindering the collection
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch06 page 286
wp
Ec
EF
recombination
Ev zone
(a) (b)
Figure 6.23 Illustration of persistent changes of (a) the density of free charge carriers in
the bulk; (b) the charge density in the space-charge region.
CIGS
+
Ec e
-
VV
CdS
energy
VV
3- EF EFn
VV
-
EF
EFp
+
h
x-coordinate x-coordinate
(a) (b)
Figure 6.24 (a) Sketch of the equilibrium band diagram of the CdS/CIGS heterojunction
involving the three charge states (VCu ,VSe )+ , (VCu ,VSe )− , and (VCu ,VSe )3− of the diva-
cancy; (b) under light or current bias, excess holes diminish the negative charge close to the
CdS/CIGS interface and excess electrons increase the effective doping in the bulk of the
material.
recombination current, and (ii) the reduction of the series resistance. Notably, these
two effects correspond to the commonly observed features during light soaking
of CIGS modules or solar cells: the increase of the open-circuit voltage and the
increase of the fill factor.
Table 6.1 Operating parameters of the best Cu(In,Ga)Se2 , CuGaSe2 and Cu(In,Ga)S2
cells.
a Confirmed total area values; b estimate. References: 1. Jackson et al. (2011); 2. Young et al. (2003);
3. Merdes et al. (2011); 4. Čhirila et al. (2013); 5. ZSW press release (2013).
Cu(In,Ga)Se 2
20
18
Efficiency η / %
16
14 Cu(In,Ga)S 2
12
10
8
1.0 1.1 1.2 1.3 1.4 1.5 1.6 1.7
Band gap energy Eg /eV
Figure 6.25 Efficiency η vs. bandgap energy E g for several series of Cu(In1−x Gax )Se2
devices with varying Ga content x (0 ≤ x ≤ 1). The plot is redrawn after Contreras et al.
(2012), and the data correspond to original data from this reference (full circles) and earlier
data from Contreras et al. (2005) (open circles), Eisenbarth et al. (2009) (open triangles),
as well as the record data from Jackson et al. (2011) (open diamonds). Additionally shown
are Cu(In,Ga)S2 results (stars) from Merdes et al. (2011).
3% below that of the best Cu(In,Ga)Se2 device is due not only to the less favourable
bandgap energy but also to the lack of the beneficial effect of small amounts of Ga
on film growth, discussed above.
The difficulty of obtaining wide-gap devices with high efficiencies is illustrated
by plotting the absorber bandgap of a series of chalcopyrite alloys vs. the attained
efficiency. Figure 6.25 shows that efficiencies in the 18–20% range peak sharply
at bandgap energies E g ∼ 1.1−1.2 eV.
One reason for the low performance of wide-gap devices is the less favourable
band offset constellation at the absorber/CdS-buffer interface. Figure 6.26 shows
the band diagram of a CuGaSe2-based heterojunction. As the increase of bandgap in
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch06 page 290
Ec
Ec Ec
EF
Ev
Ev
going from CuInSe2 to CuGaSe2 takes place almost exclusively by increase of the
energy of the conduction band, the positive band offset ∆E cab between the absorber
and the buffer in Fig. 6.17 turns into a negative one in Fig. 6.26. This implies that
p
the barrier Φb that hinders the holes from the absorber from recombining with
the electrons from the buffer does not increase proportionally with increase in the
bandgap energy. Thus the importance of interface recombination (dominated by the
p
barrier Φb ) grows considerably relative to that of bulk recombination (dominated
by E g of the absorber; see Herberholz et al., 1997b).
Using the same arguments with respect to the MoSe2 /absorber interface, the
back-surface field produced by this type of back contact turns into its contrary
because the built-in field acts in the opposite direction when the conduction-band
energy of the absorber is increased. At the moment, this drawback of Cu(In,Ga)Se2
with high Ga contents seems to be of minor importance and does not explain the
poorer performance of CuInSe2 devices either. More substantial are the changes
of the electronic quality of the bulk material with increasing x. Figure 6.27 (Hanna
et al., 2000) compares the defect density spectra of different CuIn1−x Gax Se2 alloys
with x = 0, 0.26, and 0.57.
Two features are obvious:
• The bulk defect at activation energy E a = 300 meV displays a minimum defect
density D for the composition x = 0.26. Higher bandgap material (x = 0.56),
as well as lower bandgap material (x = 0), has higher defect densities. It might
not be incidental that the superior electronic quality of the material is found at
that composition which is used to produce the highest-efficiency Cu(In,Ga)Se2
devices.
• The interface-related peak at lower activation energies for the compositions
x = 0 and 0.26 is no longer visible at x = 0.56. This is just what we would
expect from the band diagram shown in Fig. 6.28b: the Fermi level at the
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch06 page 291
16
16 9x10
10 17 5x10 16 –1 –3
–1
eV cm
–3
1x10 eV cm
–1 –3
eV cm
10 16
10 15
N2
N1 x=0 x = 0.26 x = 0.58
0.2 0.4 0.2 0.4 0.2 0.4
activation energyEa / eV
(a) (b) (c)
Figure 6.27 Admittance defect density spectra of solar cells with different CuIn1−x Gax Se2
alloy compositions. The Ga content is varied from (a) x = 0 to (c) x = 0.58. The lowest
density for the acceptor defect with an activation energy of ∼300 meV is found at (b)
x = 0.26. Source: Hanna et al. (2000).
interface moves towards mid-gap and the type inversion is no longer seen in
admittance spectroscopy.
Thus we see that bulk as well as interface properties change when going from
low Ga content Cu(In,Ga)Se2 towards higher bandgap compositions over a limit
which appears to be at x = 0.3, corresponding to a bandgap energy of 1.3 eV. A
similar limit for S/Se alloying is not yet well established. It appears, however, that
making use of the full alloy system Cu(In,Ga)(S,Se)2 enables us to go beyond the
limit of 1.3 eV while preserving a high efficiency level.
6.5.1 CuGaSe2
CuGaSe2 has a bandgap of 1.68 eV and therefore would represent an ideal partner
for CuInSe2 in an all-chalcopyrite tandem structure. However, a reasonable effi-
ciency for the top cell of any tandem structure is about 15%, far higher than has
been reached by the present polycrystalline CuGaSe2 technology. Despite intense
research during the last two decades the record efficiency is only 9.5% (Young
et al., 2003). In principle, the electronic properties of the material are not so far
from those of CuInSe2 . However, in detail, all the differences quantitatively point
in a less favourable direction. In general, the net doping density NA in CuGaSe2
appears too high (Jasenek et al., 2000). Together with the charge of deeper defects,
the high doping density leads to a strong electrical field in the space-charge region,
which enhances recombination by tunnelling.
Furthermore, a structural difference exists between the bulk CuGaSe2 (chal-
copyrite) phase and the surface (defect chalcopyrite) phase. Regardless of whether
or not this 135-phase is perfectly formed at the film surface or is a defect layer,
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch06 page 292
Ga/In
or S/Se
ratio
(a)
(b)
Figure 6.28 Band diagram of an optimised graded-gap device with an increasing Ga/In
ratio towards the back surface (region I) and an increasing S/Se-ratio towards the front
(region III). Region II has no grading. The dotted lines correspond to the conduction- and
valence-band energies of the non-graded device.
a lattice mismatch between the surface layer and the bulk material could account
for the increased defect density, which seems operative for CuGaSe2 as well as for
Cu(In,Ga)Se2 alloys with high Ga contents (Contreras et al., 1997b).
6.5.2 Cu(In,Ga)S2
The major difference between CuInS2 and Cu(In,Ga)Se2 is that the former cannot
be prepared with an overall Cu-poor composition. Cu-poor CuInS2 displays an
extremely low conductivity, making it almost unusable as a photovoltaic absorber
material (Walter et al., 1992). Even at small deviations from stoichiometry on the
In-rich side, segregation of the spinel phase is observed (Walter and Schock, 1993).
Instead, the material of choice is Cu-rich CuInS2 . As in the case of CuInSe2 , a Cu-
rich preparation route implies the removal of the unavoidable secondary phase
(here, Cu2−y S) by etching the absorber in KCN (Scheer et al., 1993). Such an
etch may involve some damage to the absorber surface as well as the introduc-
tion of shunt paths between the front and back electrodes. However, as shown
in Table 6.1, the best Cu(In,Ga)S2 device has an efficiency above 12% (Merdes
et al., 2011). Remarkably, this is achieved by a sulphurisation process rather than
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch06 page 293
p
valence-band-edge energy in Fig. 6.28 leads to an increase of the barrier Φb and
thus minimises interface recombination. The favourable combination of In/Ga-
grading towards the back contact and S/Se grading towards the front contact appears
to be an inherent feature of the combined selenisation/sulphurisation processes,
which has led to the highest efficiency devices obtained to date from sequential
processing of stacked metallic or stacked elemental precursors.
6.7 Conclusions
The objective of this chapter was not only to describe the achievements of
Cu(In,Ga)Se2 -based solar cells, but also to give an account of our present under-
standing of the physical properties of the materials involved and the electronic
behaviour of the devices. The fortunate situation of Cu(In,Ga)Se2 , which is in a
leading position among polycrystalline thin-film solar cell materials, arises from
the benign, forgiving nature of the bulk materials and interfaces. Nevertheless, we
want to draw the attention of the reader also to the work that has still to be done.
Know-how must be transferred into know-why.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch06 page 295
Three of the four cornerstones for the recent achievements mentioned in Sec-
tion 6.3 concern the growth of the films: the optimised deposition conditions, and
the incorporation of Na and Ga. However, there is still no detailed model available
definitively to describe the growth of Cu(In,Ga)Se2 , and especially the impact of
Na, which in our opinion is the most important of the different ingredients avail-
able to tune the electronic properties of the absorber. A clearer understanding of
Cu(In,Ga)Se2 growth would allow us to find optimised conditions in the wide
parameter space available, and thus to reduce the number of recombination centres
and compensating donors and optimise the number of shallow acceptors.
The deposition of the buffer layer, or more generally speaking, the forma-
tion of the heterojunction, is another critical issue. The surface chemistry during
heterojunction formation and post-deposition treatments is decisive for the final
device performance. Both processes greatly affect not only the surface defects (i.e.,
recombination and charge), and therefore the charge distribution in the device, but
also the defects in the bulk of the absorber. Concentrated effort and major progress
in these tasks would not only allow us to push the best efficiencies further above
21%, but would also provide a sound knowledge base for the various attempts at
commercialisation of Cu(In,Ga)Se2 solar cells.
References
Abken A., Heinemann F., Kampmann A., Leinkühler G., Rechid J., Sittinger V., Wietler T.
and Reineke-Koch R. (1998), ‘Large area electrodeposition of Cu(In,Ga)Se2 precursors
for the fabrication of thin film solar cells’, Proc. 2nd World Conf. on Photovoltaic Solar
Energy Conversion, Vienna, European Commission, pp. 1133–1136.
Abou-Ras D., Schorr S. and Schock H. W. (2007), ‘Grain-size distributions and grain bound-
aries of chalcopyrite-type thin films,’ J. Appl. Crystallogr. 40, 841–848.
Arora N. D., Chamberlain S. G. and Roulston D. J. (1980), ‘Diffusion length determination
in p-n junction diodes and solar cells’, Appl. Phys. Lett. 37, 325–327.
Bardeen J. (1947), ‘Surface states and rectification at a metal/semiconductor contact’, Phys.
Rev. 71, 717–727.
Binsma J. J. M. and van der Linden H. A. (1982), ‘Preparation of thin CuInS2 films via a
two-stage process’, Thin Solid Films 97, 237–243.
Birkmire R. W., McCandless B. E., Shafarman W. N. and Varrin R. D. (1989), ‘Approaches
for high efficiency CuInSe2 solar cells’, Proc. 9th Eur. Photovoltaic Solar Energy
Conf., Freiburg, Kluwer Academic Publishers, Dordrecht, pp. 134–137.
Braunger D., Hariskos D. and Schock H. W. (1998a), ‘Na-related stability issues in highly
efficient polycrystalline Cu(In,Ga)Se2 solar cells’, Proc. 2nd World Conf. on Photo-
voltaic Solar Energy Conversion, Vienna, European Commission, pp. 511–514.
Braunger D., Zweigart S. and Schock H. W. (1998b), ‘The influence of Na and Ga on
the incorporation of the chalcogen in polycrystalline Cu(In,Ga)(S,Se)2 thin-films for
photovoltaic applications’, Proc. 2nd World Conf. on Photovoltaic Solar Energy Con-
version, Vienna, European Commission, pp. 1113–1116.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch06 page 296
Brendel R. and Rau U. (1999), ‘Injection and collection diffusion lengths of polycrystalline
thin-film solar cells’, Solid State Phenomena 67–68, 81–86.
Bube R. H. (1992), Photoelectronic Properties of Semiconductors, Cambridge University
Press, Cambridge.
Bube R. H. (1998), Photovoltaic Materials, Imperial College Press, London.
Burgelman M., Engelhardt F., Guillemoles J.-F., Herberholz R., Igalson M., Klenk R.,
Lampert M., Meyer T., Nadenau V., Niemegeers A., Parisi J., Rau U., Schock H.-
W., Schmitt M., Seifert O., Walter T. and Zott S. (1997), ‘Defects in Cu(In,Ga)Se2
semiconductors and their role in the device performance of thin-film solar cells’, Progr.
Photovoltaics 5, 121–130.
Cahen D. (1987), ‘Some thoughts about defect chemistry in ternaries’, Proc. 7th Int.
Conf. on Ternary and Multinary Compounds, Mat. Res. Soc., Pittsburgh, PA,
pp. 433–442.
Cahen D. and Noufi R. (1989), ‘Defect chemical explanation for the effect of air anneal on
CdS/CuInSe2 solar cell performance’, Appl. Phys. Lett. 54, 558–560.
Čhirila A., Reinhard P., Pianezzi F., Bloesch P., Uhl A. R., Fella C., Kranz L., Keller
D., Gretener C., Hagendorfer H., Jaeger D., Erni R., Nishiwaki S., Buecheler S. and
Tiwari A. N. (2013), ‘Potassium-induced surface modification of Cu(In,Ga)Se2 thin
films for high-efficiency solar cells’, Nature Mater. 12, 1107–1111.
Chu T. L., Chu S. C., Lin S. C. and Yue J. (1984), ‘Large grain copper indium diselenide
films’, J. Electrochem. Soc. 131, 2182–2184.
Contreras M. A., Egaas B., Dippo P., Webb J., Granata J., Ramanathan K., Asher S., Swart-
zlander A. and Noufi R. (1997a), ‘On the role of Na and modifications to Cu(In,Ga)Se
absorber materials using thin MF (M = Na, K, Cs) precursor layers’, Conf. Record
26th IEEE Photovoltaic Specialists Conf., Anaheim, CA, IEEE Press, Piscataway,
pp. 359–362.
Contreras M. A., Egaas B., Ramanathan K., Hiltner J., Swartzlander A., Hasoon F. and
Noufi R. (1999), ‘Towards 20% efficiency in Cu(In,Ga)Se2 polycrystalline solar cells’,
Progr. Photovoltaics 7, 311–316.
Contreras M. A., Mansfield L. M., Egaas B., Li J., Romero M., Noufi R., Rudiger-Voigt E.
and Mannstadt W. (2012), ‘Wide bandgap Cu(In,Ga)Se2 solar cells with improved
energy conversion efficiency,’ Progr. Photovoltaics 20, 843–850.
Contreras M. A., Ramanathan K., AbuShama J., Hasoon F., Young D. L., Egaas B. and
Noufi R. (2005), ‘Diode characteristics in ZnO/CdS/Cu(In(1−x)Gax )Se2 solar cells,’
Progr. Photovoltaics 13, 209–216.
Contreras M. A., Wiesner H., Tuttle J., Ramanathan K. and Noufi R. (1997b), ‘Issues of
the chalcopyrite/defect-chalcopyrite junction model for high-efficiency Cu(In,Ga)Se2
solar cells’, Solar Energy Mater. Solar Cells 49, 239–247.
Coutts T. J., Kazmerski L. L. and Wagner S. (1986, ed.), Ternary Chalcopyrite Semicon-
ductors: Growth, Electronic Properties, and Applications, Elsevier, Amsterdam.
Devaney W. E., Chen W. S., Steward J. M. and Mickelson R. A. (1990), ‘Structure and
properties of high efficiency ZnO/CdZnS/CuInGaSe2 solar cells’, IEEE Trans. Electron
Devices ED–37, 428–433.
Dhingra A. and Rothwarf A. (1996), ‘Computer simulation and modeling of graded bandgap
CuInSe2 /CdS based solar cells’, IEEE Trans. Electron Devices 43, 613–621.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch06 page 297
Dimmler B. and Schock H. W. (1996), ‘Scaling-up of CIS technology for thin-film solar
modules’, Progr. Photovoltaics 4, 425–433.
Dullweber T., Hanna G., Shams-Kolahi W., Schwartzlander A., Contreras M. A., Noufi R.
and Schock H. W. (2000), ‘Study of the effect of gallium grading in Cu(In,Ga)Se2 ’,
Thin Solid Films 361–362, 478–481.
Eberspacher C., Pauls K. L. and Fredric C. V. (1998), ‘Improved processes for forming
CuInSe2 films’, Proc. 2nd World Conf. on Photovoltaic Solar Energy Conversion,
Vienna, European Commission, pp. 303–306.
Eisenbarth T., Unold T., Caballero R., Kaufmann C. A., Abou-Ras D. and Schock H. W.
(2009), ‘Origin of defects in CuIn1−x Gax Se2 solar cells with varied Ga content,’ Thin
Solid Films 517, 2244–2247.
Engelhardt F., Bornemann L., Köntges M., Meyer Th., Parisi J., Pschorr-Schoberer E., Hahn
B., Gebhardt W., Riedl W. and Rau U. (1999), ‘Cu(In,Ga)Se2 solar cells with a ZnSe
buffer layer — interface characterization by quantum efficiency measurements’, Progr.
Photovoltaics 7, 423–436.
Engelhardt F., Schmidt M., Meyer Th., Seifert O., Parisi J. and Rau U. (1998), ‘Metastable
electrical transport in Cu(In,Ga)Se2 thin films and ZnO/CdS/ Cu(In,Ga)Se2 het-
erostructures’, Phys. Lett. A 245, 489–493.
Friedlmeier T. M., Braunger D., Hariskos D., Kaiser M., Wanka H. N. and Schock H. W.
(1996), ‘Nucleation and growth of the CdS buffer layer on Cu(In,Ga)Se2 thin films’,
Conf. Record 25th IEEE Photovoltaic Specialists Conf., Washington D.C., IEEE Press,
Piscataway, pp. 845–848.
Gabor A. M., Tuttle J. R., Albin D. S., Contreras M. A., Noufi R. and Hermann A. M. (1994),
‘High-efficiency CuInx Ga1−x Se2 solar cells from (Inx Ga1−x )2 Se3 precursors’, Appl.
Phys. Lett. 65, 198–200.
Gabor A. M., Tuttle J. R., Bode M. H., Franz A., Tennant A. L., Contreras M. A., Noufi R.,
Jensen D. G. and Hermann A. M (1996), ‘Bandgap engineering in Cu(In,Ga)Se2 thin
films grown from (In,Ga)2 Se3 precursors’, Solar Energy Mater. Solar Cells 41, 247–
260.
Gallon P. N. R., Orsal G., Artaud M. C. and Duchemin S. (1998), ‘Studies of CuInSe2 and
CuGaSe2 thin films grown by MOCVD from three organometallic sources’, Proc. 2nd
World Conf. on Photovoltaic Solar Energy Conversion, Vienna, European Commission,
pp. 515–518.
Gilson M., Bacewicz R. and Schock H. W. (1995), “‘Dangling bonds” in CuInSe2 and related
compounds’, Proc. 13th Eur. Photovoltaic Solar Energy Conf., Nice, H. S. Stephens
& Associates, Bedford, pp. 2076–2079.
Gloeckler M., Sites J. R. and Metzger W. K. (2005), ‘Grain-boundary recombination in
Cu(In,Ga)Se2 solar cells,’ J. Appl. Phys., 98, 113704–113704-10.
Gokmen T., Gunawan O., Todorov T. K. and Mitzi D. B. (2013), ‘Band tailing and efficiency
limitation in kesterite solar cells’, Appl. Phys. Lett. 103, 103506.
Gray J. L. and Lee Y. J. (1994), ‘Numerical modeling of graded bandgap CIGS solar cells’,
Proc. 1st World Conf. on Photovoltaic Solar Energy Conversion, Waikoloa, IEEE Press,
Piscataway, NJ, pp. 123–126.
Green M. A. (1986), Solar Cells: Operating Principles, Technology and System Applica-
tions, University of New South Wales, Sydney.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch06 page 298
Green M. A. (1995), Silicon Solar Cells: Advanced Principles and Practice, University of
New South Wales, Sydney.
Green M. A. (1996a), ‘Depletion region recombination in silicon thin-film multilayer solar
cells’, Progr. Photovoltaics 4, 375–380.
Green M. A. (1996b), ‘Bounds upon grain boundary effects in minority carrier semiconduc-
tor devices: a rigorous ‘perturbation’ approach with application to silicon solar cells’,
J. Appl. Phys. 80, 1515–1521.
Haalboom T., Gödecke T., Ernst F., Rühle M., Herberholz R., Schock H.-W., Beilharz C.
and Benz K. W. (1997), ‘Phase relations and microstructure in bulk materials and thin
films of the ternary system Cu–InSe’, Inst. Phys. Conf. Ser. 152E, 249–252.
Hafemeister M., Siebentritt S., Albert J., Lux-Steiner M. C. and Sadewasser S. (2010),
‘Large neutral barrier at grain boundaries in chalcopyrite thin films,’ Phys. Rev. Lett.
104, 196602–196606.
Hahn H., Frank G., Klingler W., Meyer A. and Störger G. (1953), ‘Über einige ternäre
Chalkogenide mit Chalkopyritstruktur’, Z. anorg. u. allg. Chemie 271, 153–170.
Hanna G., Jasenek A., Rau U. and Schock H. W (2000), ‘Open circuit voltage limitations in
CuIn1−x Gax Se2 thin film solar cells — dependence on alloy compositions’, Physica
Status Solidi A. 179, R7–R8.
Hedström J., Ohlsen, H., Bodegard M., Kylner A., Stolt L., Hariskos D., Ruckh M. and
Schock H.-W. (1993), ‘ZnO/CdS/Cu(In,Ga)Se2 thin film solar cells with improved per-
formance’, Conf. Record 23rd IEEE Photovoltaic Specialists Conf., Louisville, IEEE
Press, Piscataway, NJ, pp. 364–371.
Hengel I., Neisser A., Klenk R. and Lux-Steiner C.-M. (2000), ‘Current transport in
CuInS2 :Ga/CdS/ZnO solar cells’, Thin Solid Films 361–362, 458–462.
Herberholz R. (1998), ‘Defect characterisation in chalcopyrite-based heterostructures’, Inst.
Phys. Conf. Ser. 152E, 733–740.
Herberholz R., Braunger D., Schock H. W., Haalboom T. and Ernst F. (1997a), ‘Performance
and defects in Cu(In,Ga)Se2 heterojunctions: combining electrical and structural mea-
surements’, Proc. 14th Eur. Photovoltaic Solar Energy Conf., Barcelona, H. S. Stephens
& Associates, Bedford, pp. 1246–1249.
Herberholz R., Nadenau V., Rühle U., Köble C., Schock H. W. and Dimmler B. (1997b),
‘Prospects of wide-gap chalcopyrites for thin film photovoltaic modules’, Solar Energy
Mater. Solar Cells 49, 227–237.
Herberholz R., Igalson M. and Schock H. W. (1998), ‘Distinction between bulk and inter-
face states in CuInSe2 /CdS/ZnO by space charge spectroscopy’, J. Appl. Phys. 83,
318–325.
Herberholz R., Rau U., Schock H. W., Hallboom T., Gödecke T., Ernst F., Beilharz C., Benz
K. W. and Cahen D. (1999), ‘Phase segregation, Cu migration and junction formation
in Cu(In,Ga)Se2 ’, Eur. Phys. J.: Appl. Phys. 6, 131–139.
Heske C., Fink R., Umbach E., Riedl W. and Karg F. (1996), ‘Na-induced effects on the
electronic structure and composition of Cu(In,Ga)Se2 thin-film surfaces’, Appl. Phys.
Lett. 68, 3431–3432.
Heske C., Richter G., Chen Z., Fink R., Umbach E., Riedl W. and Karg F. (1997), ‘Influence
of Na and H2 O on the surface properties of Cu(In,Ga)Se2 thin films’, J. Appl. Phys.
82, 2411–2420.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch06 page 299
Holz J., Karg F. and von Phillipsborn H. (1994), ‘The effect of substrate impurities on the
electronic conductivity in CIGS thin films, Proc. 12th Eur. Photovoltaic Solar Energy
Conf., Amsterdam, H. S. Stephens & Associates, Bedford, pp. 1592–1595.
Igalson M. and Schock H. W. (1996), ‘The metastable changes of the trap spectra of CuInSe2 -
based photovoltaic devices’, J. Appl. Phys. 80, 5765–5769.
Jackson P., Hariskos D., Lotter E., Paetel S., Wuerz R., Menner R., Wischmann W. and
Powalla M. (2011), ‘New world record efficiency for Cu(In,Ga)Se2 thin-film solar
cells beyond 20%,’ Progr. Photovoltaics 19, 894–897.
Jäger-Waldau G., Schmid D. and Jäger-Waldau A. (1991), ‘Diffusion length measurement
of heterojunction thin films by junction-EBIC’, J. Phys. IV 1, C6, 131–132.
Jasenek A., Rau U., Nadenau V. and Schock H. W. (2000), ‘Electronic properties of
CuGaSe2 -based heterojunction solar cells. Part II — Defect spectroscopy’, J. Appl.
Phys. 87, 594–602.
Jensen N., Rau U., Hausner R. M., Uppal S., Oberbeck L., Bergmann R. B. and Werner J. H.
(2000), ‘Recombination mechanisms in amorphous silicon/crystalline silicon hetero-
junction solar cells’, J. Appl. Phys. 87, 2639–2645.
Kapur V. K., Basol B. M. and Tseng E. S. (1987), ‘Low-cost methods for the production of
semiconductor films for CuInSe2 /CdS solar cells’, Solar Cells 21, 65–72.
Kazmerski L. L. and Wagner S. (1985), ‘Cu-ternary chalcopyrite solar cells’, in Coutts T. J.
and Meakin J. D., Current Topics in Photovoltaics, Academic Press, Orlando.
Kessler J., Velthaus K. O., Ruckh M., Laichinger R., Schock H. W., Lincot D., Ortega R.
and Vedel J. (1992), ‘Chemical bath deposition of CdS on CuInSe2 , etching effects and
growth kinetics’, Proc. 6th Int. Photovoltaic Solar Energy Conf., New Delhi, Oxford
IBH Publishing, New Delhi, pp. 1005–1010.
Keyes B. M., Hasoon F., Dippo P., Balcioglu A. and Aboulfotuh F. (1997), ‘Influence of Na on
the electro-optical properties of Cu(In,Ga)Se2 ’, Conf. Record 26th IEEE Photovoltaic
Specialists Conf., Anaheim, IEEE Press, Piscataway, NJ, pp. 479–482.
Klein A. and Jaegermann W. (1996), ‘Chemical interaction of Na-cleaved (001) surfaces of
CuInSe2 ’, J. Appl. Phys. 80, 5039–5043.
Klenk R. and Schock H. W. (1994), ‘Photocurrent collection in thin film solar cells —
calculation and characterization for CuGaSe2 /(Zn,Cd)S’, Proc. 12th Eur. Photovoltaic
Solar Energy Conf., Amsterdam, H. S. Stephens & Associates, Bedford, pp.1588–1591.
Klenk R., Walter T., Schock H. W. and Cahen D. (1993), ‘A model for the successful growth
of polycrystalline films of CuInSe2 by multisource physical vacuum evaporation’, Adv.
Mat. 5, 114–119.
Kronik L., Cahen D. and Schock H. W. (1998), ‘Effects of sodium on polycrystalline
Cu(In,Ga)Se2 and its solar cell performance’, Adv. Mater. 10, 31–36.
Kronik L., Rau U., Guillemoles J.-F., Braunger D., Schock H. W. and Cahen D. (2000),
‘Interface redox engineering of Cu(In,Ga)Se2 -based solar cells: oxygen, sodium, and
chemical bath effects’, Thin Solid Films 361–362, 353–359.
Kushiya K. (2013), ‘CIS-based thin-film PV technology in solar frontier K.K.’, Solar Energy
Mater. Solar Cells 122, 309–313.
Kushiya K., Tachiyuli M., Nagoya Y., Fujimaki A., Sang B., Okumara D., Satoh M. and
Yamase O. (1999), ‘Progress in large-area Cu(In,Ga)Se2 -based thin-film modules with
a Zn(O,S,OH) x buffer layer’, Tech. Digest 11th Int. Sci. Eng. Conf., Tokyo University
of Agriculture and Technology, Tokyo, Japan, pp. 637–640.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch06 page 300
Nelson A. J., Schwerdtfeger C. R., Wei S.-H., Zunger A., Rioux D., Patel R. and Höchst H.
(1993), ‘Theoretical and experimental studies of the ZnSe/CuInSe2 heterojunction
band offset’, Appl. Phys. Lett. 62, 2557–2559.
Neumann H. (1986), ‘Optical properties and electronic band structure of CuInSe2 ’, Solar
Cells 16, 317–333.
Niemegeers A. and Burgelman M. (1996), ‘Numerical modelling of ac-characteristics of
CdTe and CIS solar cells’, Proc. 25nd IEEE Photovoltaic Specialists Conf., Washington
D.C., April 1996, IEEE, New York, NY, pp. 901–904.
Niemegeers A., Burgelman M. and de Vos A. (1995), ‘On the CdS/CuInSe2 conduction
band discontinuity’, Appl. Phys. Lett. 67, 843–845.
Niemegeers A., Burgelman M., Herberholz R., Rau U., Hariskos D. and Schock H. W.
(1998), ‘Model for electronic transport in Cu(In,Ga)Se2 solar cells’, Progr. Photo-
voltaics 6, 407–421.
Niki S., Fons P. J., Yamada A., Suzuki R., Ohdaira T., Ishibashi S. and Oyanagai H. (1994),
‘High quality CuInSe2 epitaxial films — molecular beam epitaxial growth and intrinsic
properties’, Inst. Phys Conf. Ser. 152E, 221–227.
Niles D. W., Al-Jassim M. and Ramanathan K. (1999), ‘Direct observation of Na and
O impurities at grain surfaces of CuInSe2 thin films’, J. Vac. Sci. Technol. A 17,
291–296.
Niles D. W., Ramanathan K., Haason F., Noufi R., Tielsh B. J. and Fulghum J. E. (1997),
‘Na impurity chemistry in photovoltaic CIGS thin films: investigation with X-ray pho-
toelectron spectroscopy’, J. Vac. Sci. Technol. A 15, 3044–3049.
Noufi R., Axton R., Herrington C. and Deb S. K. (1984), ‘Electronic properties versus
composition of thin films of CuInSe2 ’, Appl. Phys. Lett. 45, 668–670.
Padovani F. A. and Stratton R. (1966), ‘Field and thermionic field emission in Schottky
barriers’, Solid State Electronics 9, 695–707.
Parisi J., Hilburger D., Schmitt M. and Rau U. (1998), ‘Quantum efficiency and admittance
spectroscopy on Cu(In,Ga)Se2 solar cells’, Solar Energy Mater. Solar Cells 50, 79–85.
Persson C. and Zunger A. (2003), ‘Anomalous grain boundary physics in polycrystalline
CuInSe2 : the existence of a hole barrier,’ Phys. Rev. Lett., 91, 266401–266405.
Polizzotti A., Repins I. L., Noufi R., Wei S.-H. and Mitzi D. B. (2013), ‘The state and future
prospects of kesterite photovoltaics’, Energy Environ. Sci. 6, 3171–3182.
Potter R. R., Eberspacher C. and Fabick L. B. (1985), ‘Device analysis of CuInSe2 / (Cd,Zn)S
solar cells’, Conf. Record 18th IEEE Photovoltaic Specialists Conf., Las Vegas, NV,
IEEE Press, Piscataway, pp. 1659–1664.
Probst V., Karg F., Rimmasch J., Riedl W., Stetter W., Harms H. and Eibl O. (1996),
‘Advanced stacked elemental layer progress for Cu(InGa)Se2 thin film photovoltaic
devices’, Mat. Res. Soc. Symp. Proc. 426, 165–176.
Ramanathan K., Bhattacharya R. N., Granata J., Webb J., Niles D., Contreras M. A.,
Wiesner H., Haason F. S. and Noufi R. (1998a), ‘Advances in CIGS solar cell research
at NREL’, Conf. Record 26th IEEE Photovoltaic Specialists Conf., Anaheim, CA, IEEE
Press, Piscataway, NJ, pp. 319–325.
Ramanathan K., Wiesner H., Asher S., Niles D., Bhattacharya R. N., Keane J., Contreras
M. A. and Noufi R. (1998b), ‘High efficiency Cu(In,Ga)Se2 thin film solar cells with-
out intermediate buffer layers’, Proc. 2nd World Conf. on Photovoltaic Solar Energy
Conversion, Vienna, European Commission, pp. 477–481.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch06 page 302
Scheer R., Walter T., Schock H. W., Fearheiley M. L. and Lewerenz H. J. (1993), ‘CuInS2
based thin film solar cells with 10.2% efficiency’, Appl. Phys. Lett. 63, 3294–3296.
Schmid D., Ruckh M., Grunwald F. and Schock H. W. (1993), ‘Chalcopyrite/defect chal-
copyrite heterojunctions on basis of CuInSe2 ’, J. Appl. Phys. 73, 2902–2909.
Schmid D., Ruckh M., Grunwald F. and Schock H. W. (1996a), ‘Photoemission stud-
ies on Cu(In,Ga)Se2 thin films and related binary selenides’, Appl. Surf. Sci. 103,
409–429.
Schmid D., Ruckh M. and Schock H. W. (1996b), ‘A comprehensive characterization of the
interfaces in Mo/CIS/CdS/ZnO solar cell structures’, Solar Energy Mater. Solar Cells
41, 281–294.
Schmidt M., Braunger D., Schäffler R., Schock H. W. and Rau U. (2000), ‘Influence of
damp heat on the electrical properties of Cu(In,Ga)Se2 solar cells’, Thin Solid Films
361–362, 283–287.
Schmitt M., Rau U. and Parisi J. (1995), ‘Investigation of deep trap levels in CuInSe2 solar
cells by temperature dependent admittance measurements’, Proc. 13th Eur. Photo-
voltaic Solar Energy Conf., Nice, H. S. Stephens & Associates, Bedford, pp. 1969–
1972.
Schock H. W. (1996), ‘Thin film photovoltaics’, Appl. Surf. Sci. 92, 606–616.
Schroeder D. J. and Rockett A. A. (1997), ‘Electronic effects of sodium in epitaxial
CuIn1−x Gax Se2 ’, J. Appl. Phys. 82, 5982–5985.
Seifert O., Engelhardt F., Meyer Th., Hirsch M. T., Parisi J., Beilharz C., Schmitt M. and
Rau U. (1997), ‘Observation of a metastability in the DC and AC electrical transport
properties of Cu(In,Ga)Se2 single crystals, thin films and solar cells’, Inst. Phys. Conf.
Ser. 152E, 253–256.
Shay J. L. and Wernick J. H. (1975), Ternary Chalcopyrite Semiconductors: Growth, Elec-
tronic Properties, and Applications, Pergamon Press, Oxford.
Siebentritt S., Igalson M., Persson C. and Lany S. (2010), ‘The electronic structure of
chalcopyrites-bands, point defects and grain boundaries,’ Progr. Photovoltaics 18, 390–
410.
Siebentritt S., Sadewasser S., Wimmer M., Leendertz C., Eisenbarth T. and Lux-Steiner
M. C. (2006), ‘Evidence for a neutral grain-boundary barrier in chalcopyrites’, Phys.
Rev. Lett. 97, 146601–146605.
Staebler D. L. and Wronski C. R. (1977), ‘Reversible conductivity changes in discharge-
produced amorphous Si’, Appl. Phys. Lett. 31, 292–294.
Stolt L., Hedström J., Kessler J., Ruckh M., Velthaus K. O. and Schock H. W. (1993),
‘ZnO/CdS/CuInSe2 thin-film solar cells with improved performance’, Appl. Phys. Lett.
62, 597–599.
Suryawanshi M. P., Agawane G. L., Bhosale S. M., Shin S. W., Patil P. S., Kim J. H. and
Moholkar A. V. (2013), ‘CZTS based thin film solar cells: a status review’, Mater.
Technol. 28, 98–109.
Takei R., Tanino H., Chichibu S. and Nakanishi H. (1996), ‘Depth profiles of spatially-
resolved Raman spectra of a CuInSe2 -based thin-film solar cell’, J. Appl. Phys. 79,
2793–2795.
Taretto K. and Rau U. (2008), ‘Numerical simulation of carrier collection and recombi-
nation at grain boundaries in Cu(In,Ga)Se2 solar cells,’ J. Appl. Phys. 103, 094523–
094523-11.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch06 page 304
Topic M., Smole F. and Furlan J. (1997), ‘Examination of blocking current–voltage behavior
through defect chalcopyrite layer in ZnO/CdS/Cu(In,Ga)Se2 solar cell’, Solar Energy
Mater. Solar Cells, 311–317.
Urbaniak A. and Igalson M. (2009), ‘Creation and relaxation of light- and bias-induced
metastabilities in Cu(In,Ga)Se2 ’, J. Appl. Phys. 106, 063720.
Wada T., Hayashi S., Hashimoto Y., Nishiwaki S., Sato T., Negami T. and Nishitani M.
(1998), ‘High efficiency Cu(In,Ga)Se2 (CIGS) solar cells with improved CIGS sur-
face’, Proc. 2nd World Conf. on Photovoltaic Solar Energy Conversion, Vienna, Euro-
pean Commission, pp. 403–408.
Wada T., Kohara N., Negami T. and Nishitani M. (1996), ‘Chemical and structural char-
acterization of Cu(In,Ga)Se2 /Mo Interface in Cu(In,Ga)Se2 solar cells’, Jpn. J. Appl.
Phys. 35, 1253–1256.
Wagner S., Shay J. L., Migliorato P. and Kasper H. M. (1974), ‘CuInSe2 /CdS heterojunction
photovoltaic detectors’, Appl. Phys. Lett. 25, 434–435.
Wallin E., Malm U., Jarmar T., Lundberg O., Edoff M. and Stolt L. (2012), ‘World-record
Cu(In,Ga)Se2 -based thin-film sub-module with 17.4% efficiency,’ Progr. Photovoltaics
20, 851–854.
Walter T., Braunger D., Hariskos D., Köble C. and Schock H. W. (1995), ‘CuInS2 : film
growth, devices, submodules and perspectives’, Proc. 13th Eur. Photovoltaic Solar
Energy Conf., Nice, H. S. Stephens & Associates, Bedford, pp. 597–600.
Walter T., Content A., Velthaus K. O. and Schock H. W. (1992), ‘Solar cells based on
CuIn(Se,S)2 ’, Solar Energy Mater. Solar Cells 26, 357–368.
Walter T., Herberholz R., Müller C. and Schock H. W. (1996a), ‘Determination of defect
distributions from admittance measurements and application to Cu(In,Ga)Se2 based
heterojunctions’, J. Appl. Phys. 80, 4411–4420.
Walter T., Herberholz R. and Schock H. W. (1996b), ‘Distribution of defects in polycrys-
talline thin films’, Solid State Phenomena 51, 309–316.
Walter T. and Schock H. W. (1993), ‘Structural and electrical investigations of the anion
exchange in polycrystalline CuIn(S,Se)2 thin films’, Jpn. J. Appl. Phys. 32–3, 116–119.
Wang W., Winkler M. T., Gunawan O., Gokmen T., Todorov T. K., Zhu Y. and
Mitzi D. B., ‘Device characteristics of CZTSSe thin-film solar cells with 12.6%
efficiency’, Advanced Energy Mater., published online: 27 Nov. 2013, DOI:
10.1002/aenm.201301465.
Weber A., Mainz R. and Schock H. W. (2010), ‘On the Sn loss from thin films of the material
system Cu–Zn–Sn–S in high vacuum’, J. Appl. Phys. 107, 013516–013516.
Wei S. H. and Zunger A. (1993), ‘Band offsets at the CdS/CuInSe2 heterojunction’, Appl.
Phys. Lett. 63, 2549–2551.
Wolf D., Müller G., Stetter W. and Karg F. (1998), ‘In-situ investigation of Cu–In–Se
reactions: impact of Na on CIS formation’, Proc. 2nd World Conf. on Photovoltaic
Solar Energy Conversion, Vienna, European Commission, pp. 2426–2429.
Würz R., Eicke A., Kessler F., Paetel S., Efimenko S. and Schlegel C. (2012), ‘CIGS thin-
film solar cells and modules on enamelled steel substrates’, Solar Energy Mater. Solar
Cells 100, 132–137.
Yan Y., Jones K. M., Jiang C. S., Wu X. Z., Noufi R. and Al-Jassim M. M. (2007), ‘Under-
standing the defect physics in polycrystalline photovoltaic materials,’ Physica B 401,
25–32.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch06 page 305
Young D. L., Keane J., Duda A., AbuShama J. A. M., Perkins C. L., Romero M. and
Noufi R. (2003), ‘Improved performance in ZnO/CdS/CuGaSe2 thin-film solar cells’,
Progr. Photovoltaics 11, 535–541.
Zhang S. B., Wei S. H. and Zunger A. (1997), ‘Stabilization of ternary compounds via
ordered arrays of defect pairs’, Phys. Rev. Lett. 78, 4059–4062.
Zhang S. B., Wei S. H., Zunger A. and Katayama-Yoshiba H. (1998), ‘Defect physics of the
CuInSe2 chalcopyrite semiconductor’, Phys. Rev B. 57, 9642–9656.
Zweibel K. (1995), ‘Thin films: past, present, future’, Progr. Photovoltaics 3, 279–293.
May 2, 2013 14:6 BC: 8831 - Probability and Statistical Theory PST˙ws
CHAPTER 7
SUPER-HIGH-EFFICIENCY III–V
TANDEM AND MULTIJUNCTION CELLS
MASAFUMI YAMAGUCHI
Toyota Technological Institute, Nagoya 468–8511, Japan
[email protected]
The reasonable man adapts himself to the world: the unreasonable one persists in trying
to adapt the world to himself. Therefore all progress depends on the unreasonable man.
George Bernard Shaw, Man and Superman, 1903.
7.1 Introduction
Although solar electricity, including solar photovoltaics, is expected to make a
great contribution as a major energy source, providing a share of about 20% of
global electric power in 2050 and about 70% in 2100 (WBGU, 2003), nuclear
power is still a major energy source because of its huge power-generation capacity
and relatively low electricity cost. In order to realise the vision of a solar electricity
future, high-performance solar cells are very attractive.
The development of high-performance solar cells offers a promising path-
way toward achieving high power per unit cost for many applications. Substantial
increases in conversion efficiency can be realised by using multijunction solar cells
rather than single-junction cells. The principles of multijunction cells were sug-
gested as long ago as 1955 (Jackson, 1955), and they have been investigated since
1960 (Wolf, 1960), as shown in the historical timeline of Table 7.1. However, no
significant progress was made in multijunction cell conversion efficiency during
the period 1960–75 because of poor thin-film fabrication technologies. It is thanks
to progress in the technology of liquid-phase epitaxy (LPE) and vapour-phase
epitaxy and to optical devices such as semiconductor lasers that more efficient
AlGaAs/GaAs tandem cells were developed during the 1980s. These included tun-
nel junctions (Hutchby et al., 1985) and metal interconnections (Ludowise et al.,
1982; Flores, 1983; Chung et al., 1989).
At that time, the predicted efficiency of close to 30% (Fan et al., 1982) was
not achieved because of difficulties in making high-performance, stable tunnel
307
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch07 page 308
308 M. Yamaguchi
Location
Date Milestone or individual
junctions (Yamaguchi et al., 1987), and also because of the presence of oxygen-
related defects in the AlGaAs (Ando et al., 1987). High-performance, stable tunnel
junctions with a double-hetero (DH) structure (in which the GaAs tunnel junction
is sandwiched between AlGaAs layers) were developed by Sugiura et al. (1988)
of NTT Electrical Communications Laboratories (NTT). The use of InGaP for
the top cell was introduced by Olson et al. (1990) of National Renewable Energy
Laboratory (NREL), and as a result, a GaInP/GaAs tandem cell with an efficiency
of 29.5%, though a small area of only 0.25 cm2 , was made by Bertness et al.
(1994). Monolithically grown InGaP/GaAs two-junction solar cells achieved a
then-record efficiency of 30.3% in 1997 (Takamoto et al., 1997a) at 1 Sun AM
1.5. As regards concentrator systems, over 30% efficiency was attained by the
mechanically stacked GaAs/GaSb cells of Fraas et al. (1990).
InGaP/GaAs-based multijunction (MJ) solar cells have drawn increased atten-
tion for space applications because of the superior radiation resistance of InGaP
top cells and materials, which was discovered by the author and co-workers (Yam-
aguchi et al., 1997). In addition we also achieved the possibility of high conversion
efficiency of over 30%. The commercial satellite HS 601HP with two-junction
GaInP/GaAs on Ge solar arrays was launched in 1997 (Brown et al., 1997).
InGaP/GaAs-based MJ solar cells have also drawn increased attention for
terrestrial applications because operating MJ cells under concentrated sun-
light has great potential for providing high-performance, low-cost solar mod-
ules. For concentrator applications, the cell contact grid structure should be
designed so as to reduce the energy loss due to series resistance; in this way
38.9% (AM 1.5G, 489 Suns) efficiency was demonstrated by Sharp (Takamoto
et al., 2005). The achievement of 41.1% efficiency under 454 Suns with
In0.65 Ga0.35 P/In0.17 GaAs0.83 /Ge three-junction concentrator cells by Fraunhofer
ISE (Bett et al., 2009) and 41.6% efficiency under 364 Suns with lattice-matched
InGaP/InGaAs/Ge three-junction concentrator cells by Spectrolab (King et al.,
2009) has also been reported. Later, 42.3% efficiency under 406 Suns with bifa-
cial epitaxially grown InGaP/GaAs/InGaAs three-junction concentrator cells was
reported by Spire (Wojtczuk et al., 2010); and Solar Junction (Solar Junction, 2011;
Sabnis, 2012) achieved 43.5% efficiency under 418 Suns with lattice-matched
InGaP/GaAs/GaInNAs three-junction concentrator cells. Most recently, 37.9%
efficiency under 1 Sun and 44.4% efficiency under 240–300 Suns have been demon-
strated with InGaP/GaAs/InGaAs three-junction cells by Sharp (Sharp, 2013;
Sasaki, 2013; Yamaguchi and Luque, 2013; Green et al., 2013). In addition, high-
efficiency large-area (5445 cm2) concentrator InGaP/InGaAs/Ge three-junction
modules for 500 Suns used with an outdoor efficiency of 31.5% (Araki et al., 2005)
have been developed as a result of incorporating high-efficiency InGaP/InGaAs/Ge
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch07 page 310
310 M. Yamaguchi
Figure 7.1 Principle of wide photoresponse by using a multijunction solar cell, for the case
of an InGaP/InGaAs/Ge triple-junction solar cell.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch07 page 311
Figure 7.2 Current–voltage curve and spectral response for an InGaP/InGaAs/Ge mono-
lithic, two-terminal three-junction cell (Takamoto et al., 2000).
the less energetic photons to the cells below. The next cell in the stack (here the
GaAs middle cell) absorbs all the transmitted photons with energies equal to or
greater than its bandgap energy, and transmits the rest downward in the stack (in this
example, to the Ge bottom cell). As shown in Fig. 7.2, the current–voltage curve and
spectral response for an InGaP/GaAs/Ge monolithic, two-terminal triple-junction
cell (Takamoto et al., 2000) shows the wideband photoresponse of MJ cells. In
principle, any number of cells can be used in tandem.
Computer analysis of the performance of MJ solar cells has been carried out by
several researchers (Loferski, 1976; Lamorte and Abbott, 1980; Mitchell, 1981; Fan
et al., 1982; Nell and Barnett, 1987; Amano et al., 1989; Letay and Bett, 2001).
The following explanations are based on the findings of Letay and Bett (2001).
Figure 7.3 shows their AM 1.5 iso-efficiency plots for a three-cell, two-terminal
tandem structure with Ge as the bottom cell at 25◦ C and 1 Sun. The maximum
theoretical efficiency for this system is 38% at AM 1.5. For optimal efficiencies in
the two-terminal structure, the allowable range of bandgaps for the top and bottom
cells is very narrow. The top cell must have a bandgap of about 1.8 eV, and the
middle cell about 1.1 eV. In this case, one of the candidate material combinations
is InGaP/InGaAs/Ge. However, because the optimal bandgap combination, which
is 1.8 eV/1.1 eV/0.66 eV, is a lattice-mismatched system, 1.85 eV/1.4 eV/0.66 eV
and 2 eV/1.4 eV/0.66 eV lattice-matched systems have been realistically devel-
oped. The optimal efficiencies of such systems are 31–34.5%. In these cases,
InGaP/GaAs/Ge, AlInGaP/GaAs/Ge and AlGaAs/GaAs/Ge are candidate material
combinations.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch07 page 312
312 M. Yamaguchi
Figure 7.3 Calculated AM 1.5 iso-efficiency plots for a three-cell, two-terminal tandem
structure with Ge as the bottom cell at 25◦ C and 1 Sun. From Letay and Bett (2001).
2.00
Three cells
Bottom-cell band gap = 1.00 eV
Series-connected
AM1
Middle-cell band gap (eV)
27 C
1.75
1.50
41.1%
39.0%
37.0%
35.0%
1.25
1.75 2.00 2.25 2.50
Top-cell band gap (eV)
Figure 7.4 Calculated AM 1 iso-efficiency plots at 27◦ C and 1 Sun for the three-cell tandem
structure with the cells connected in series. The bottom cell has a fixed energy gap of 1.0 eV
(Fan et al., 1982).
Figure 7.4 shows the AM 1 iso-efficiency plots at 27◦C for the three-cell
tandem structure with the cells connected in series (Fan et al., 1982). The curves are
plotted for a bottom-cell bandgap of 1.0 eV, because the maximum calculated effi-
ciencies are obtained for a range of 0.95–1.0 eV. The optimal top/middle/bottom
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch07 page 313
60
50
conversion efficiency / %
40
theory (Conc.)
30 theory (1 Sun)
realised (Conc.)
realised (1 Sun)
20
1 2 3 4 5 6
number of junctions
Figure 7.5 Theoretical conversion efficiencies of single-junction and multijunction solar
cells in comparison with experimentally realised efficiencies for 1 Sun intensity and under
concentration.
314 M. Yamaguchi
40
x 0.1 Factory
Equipment
30 Labour
Cost(US$/Wp)
Others
Dopant
20
V source
V' source
x 0 .1
10 Substrate
0
Present Mass MOCVD MBE CBE CBE +
production (future) (future) (future) concentrator
Figure 7.6 Cost estimates for monolithically integrated two-junction solar cells consisting
of a III–V compound solar cell combined with a Si cell, fabricated by metal-organic chemical
vapour deposition (MOCVD), molecular beam epitaxy (MBE) or chemical beam epitaxy
(CBE) (Yamaguchi et al., 1994).
The use of concentrating systems can further enhance the cost advantage of
high-efficiency cells. Figure 7.6 shows cost estimates for monolithically integrated
two-junction solar cells with a III–V compound solar cell and a Si cell, fabricated
by metal–organic chemical vapour deposition (MOCVD), molecular beam epitaxy
(MBE) or chemical beam epitaxy (CBE) (Yamaguchi et al., 1994). Through source
material cost reduction combined with the use of the tandem structure and a con-
centrator system, high-efficiency cells costing less than $1/Wp should be possible.
Concentrator operation is very effective for cost reduction of solar cell modules
and thus also of PV systems. Figure 7.7 shows a summary of the estimated cost for
the concentrator PV systems vs. concentration ratio (Yamaguchi, 2003). As shown
in Fig. 7.7, concentrator PV systems using MJ solar cells have great potential for
cost reduction to under 50 ¢/Wp if one could fabricate 35%-efficient modules for
5000 Suns use.
The effectiveness of high-performance, low-cost concentrator PV systems is
also discussed in Chapter 11 of this book by Luque-Heredia and Luque, and sep-
arately by Yamaguchi and Luque (1999).
Figure 7.7 Summary of estimated cost for the concentrator PV systems vs. concentration
ratio (Yamaguchi, 2003).
30
Laboratory
conversion efficiency (%)
25 MOCVD
MBE
20
LPE LPE
MOCV
15
10 Manufactured
(AM0)
5
0
1970 1980 1990
year
Figure 7.8 Chronological improvements in the efficiencies of GaAs solar cells fabricated
by the LPE, MOCVD and MBE methods.
fabricate GaAs solar cells in 1972 because it produces high-quality epitaxial film
and has a simple growth system. However, it is not as useful for devices that
involve multilayers because of the difficulty of controlling layer thickness, doping,
composition and speed of throughput. Since 1977, MOCVD has been used to
fabricate large-area GaAs solar cells because it is capable of large-scale, large-area
production and has good reproducibility and controllability.
Table 7.2 compares the advantages and disadvantages of the various epitax-
ial technologies. While LPE can produce high-quality epitaxial films, MOCVD is
effective for the large-scale, large-area production of solar cells. Molecular beam
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch07 page 316
316 M. Yamaguchi
+ + + +
p+ p+ p+ p+
top cell
metal
n n interconnect n – n
interconnect
tunnel
++ +
n n + –
++ + n
p p + +
p p n p
bottom cell
n n p n
– – + –
epitaxy and CBE are advantageous for realising novel multilayer structures such as
MJ solar cells because they provide excellent controllability of monolayer abrupt-
ness and thickness due to the nature of the beam (Yamaguchi et al., 1994). However,
so far there have been few reports of CBE-grown solar cells.
structure, only one external circuit load is needed, but the photocurrents in the
two cells must be equal for optimal operation. Key issues for maximum-efficiency
monolithic cascade cells (two-terminal MJ cells series connected with tunnel junc-
tion) are the formation of tunnel junctions of high performance and stability for cell
interconnection, and the growth of optimum bandgap top- and bottom-cell struc-
tures on lattice-mismatched substrates, without permitting propagation of delete-
rious misfit and thermal stress-induced dislocations.
In contrast, the photocurrents in three-and four-terminal cells do not have to
be equal. However, it is very difficult to connect three-terminal devices in series, so
three-terminal tandem cells do not appear to be viable. In the four-terminal case,
two separate external circuit loads are used. Since the two individual cells are not
coupled, the photocurrents do not have to be the same. Consequently, a much larger
range of bandgap energy combinations is possible, and the changes in photocurrents
with changing solar spectral distributions do not pose serious limits. This approach
avoids the problems of lattice-mismatched epitaxial growth, current matching and
the internal electrical connection of the two-terminal device. Important issues for
obtaining high-efficiency mechanically stacked cells are the development of MJ
cell fabrication techniques such as thinning the top cell, bonding the bottom cell
to the top cell, and cell connections.
τ = 1/B N (7.1)
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch07 page 318
318 M. Yamaguchi
30
25
20
efficiency / %
10
0
0.1 1 10 100
minority-carrier diffusion length / μm
Figure 7.10 Minority-carrier diffusion length dependence of GaAs single-junction solar
cell efficiency.
1000
S=0 (cm/s)
Relative PL Intensity (Arb. Unit)
5800
Growth Temp. (°C)
100 104
700
650 105
750
10 106
Without
Buffer Layer 3x106
1
Without
0.1
AlInP Window
0.01
0.01 0.1 1 10 100
Minority Carrier Lifetime τ (ns)
Figure 7.12 Changes in photoluminescence (PL) intensity of the solar cell active layer as a
function of the minority-carrier lifetime and surface recombination velocity in InGaP, grown
by MOCVD.
et al., 1987) that an oxygen-related defect in the AlGaAs top-cell materials acts as
a recombination centre. As a top-cell material lattice-matched to GaAs or Ge sub-
strates, InGaP has some advantages (Olson et al., 1990), such as a lower interface
recombination velocity, a lower oxygen-related defect problem and a good window
layer material compared with AlGaAs. The top-cell characteristics depend on the
minority-carrier lifetime in the top-cell layers.
Figure 7.12 shows changes in photoluminescence (PL) intensity of the solar
cell active layer as a function of the minority-carrier lifetime τ of the p-InGaP base
layer grown by MOCVD and surface recombination velocity S. The lowest level of
S was obtained by introducing an AlInP window layer, and the highest minority-
carrier lifetime τ was obtained by introducing a buffer layer and optimising the
growth temperature. The best conversion efficiency of the InGaP single-junction
cell was 18.5% (Yang et al., 1997).
320 M. Yamaguchi
lifetime τ was calculated from the following equation (Yamaguchi and Amano,
1985):
35
S=1E+4 cm/s
–2
S=1E+5 cm/s
30
short-circuit current density / A cm
S=1E+6 cm/s
S=1E+7 cm/s
25
20
15
10
0
0 0.5 1 1.5 2 2.5
junction depth / μm
Figure 7.14 Surface recombination effect velocity on the short-circuit current density of a
In0.14 Ga0.86 As homojunction solar cell as a function of junction depth.
than 50 nm is necessary. In order to decrease the efficiency drop due to front and
rear surface recombination as shown in Fig. 7.15, formation of a heteroface or
double-hetero structure is necessary.
Figure 7.15 shows the changes in Voc and i sc of InGaP single-junction cells as
a function of the potential barrier E. A wide-bandgap back-surface field (BSF)
layer (Takamoto et al., 2003) is found to be the most effective for confinement of
minority carriers.
322 M. Yamaguchi
Figure 7.15 Change in Voc and i sc of InGaP single-junction cells as a function of the
potential barrier E.
in Fig. 7.17, a double-hetero (DH) structure was found by the authors to be useful
for preventing diffusion (Sugiura et al., 1988). An InGaP tunnel junction was tried
for the first time for an InGaP/GaAs tandem cell in our work (Takamotoet al.,
1997b). Zn and Si were used as p-type and n-type dopants, respectively. The peak
tunnelling current of the InGaP tunnel junction increased from 5 mA cm−2 up to
2 A cm−2 on making a DH structure with AlInP barriers. Effective suppression of
the Zn diffusion from tunnel junction by the InGaP tunnel junction with the AlInP-
DH structure is thought to be due to the lower diffusion coefficient (Takamoto
et al., 1999) for Zn in the wider bandgap energy materials such as the AlInP barrier
layer and InGaP tunnel junction layer. In conclusion, the InGaP tunnel junction is
very effective for obtaining high tunnelling currents, and the DH structure is useful
for preventing diffusion.
Table 7.3 summarises factors of importance in achieving MJ cells of very high
efficiency.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch07 page 323
104
tunnel peak current density / A cm
3
10
–2
isc of bottom cell / mA cm
2 15
10
10
1 10
10–1
10–2 5
10–3
–4
10 0
0.5 1.0 1.5 2.0
bandgap energy / eV
Figure 7.16 Calculated tunnel peak current density and short-circuit current density of a
GaAs bottom cell as a function of the bandgap energy of the tunnel junction.
10
–2
tunnel peak current density / A cm
10–1
–2
10
–3
10
10
500 600 700 800
o
annealing temperature / C
Figure 7.17 Annealing temperature dependence of tunnel peak current densities for double-
hetero structure tunnel diodes. X is the Al mole fraction in the Al x Ga1−x As barrier layers.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch07 page 324
324 M. Yamaguchi
Table 7.3 Key issues for realising super-high-efficiency multijunction solar cells.
Figure 7.18 Schematic illustration of a triple-junction cell and approaches for improving
its efficiency.
326 M. Yamaguchi
low diffusion coefficients, are used as dopants for p-type AlGaAs and n-type
(Al)InGaP, respectively. Furthermore, the double-hetero structure appears to sup-
press impurity diffusion from the highly doped tunnel junction (Takamoto et al.,
1999). The second tunnel junction between the middle and bottom cells consists
of p-InGaP/ p-(In)GaAs/n-(In)GaAs/n-InGaP, which has a wider bandgap than the
middle cell materials.
Figure 7.19 Change in the spectral response due to modification of the first heterolayer
from GaAs to InGaP (without antireflection coating).
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch07 page 327
0% In 1% In 2% In
100 µm
Figure 7.20 Surface morphology of InGaAs with various indium compositions grown
on Ge.
328 M. Yamaguchi
600
power density / W m–2
500
400
300
200
100
0
10 15 20 25 30 35 40
BOL efficiency / %
Figure 7.21 Effectiveness of radiation resistance and high conversion efficiency of space
cells in increasing the power density of space missions. Beginning-of-life efficiency, BOL,
is the efficiency of the space solar cells before space satellite launching.
0.8
15
1 MeV - 10 e / cm –2
0.78 photo-injection: 100 mW / cm
75°C
power ratio PI / P0
0.76
50°C
0.74 25°C
0.72
0.7
0 10 20 30 40
photo-injection time / min
Figure 7.22 Maximum power recovery of the InGaP/GaAs tandem cell due to light illu-
mination at various temperatures.
Figure 7.23 Deep-level transient spectroscopy (DLTS) spectrum of trap H2 (Ev + 0.55 eV)
for various injection times at 25◦ C with an injection density of 100 mA cm−2 .
330 M. Yamaguchi
InGaP/GaAs-3J
efficiency / %
silicon
year
Figure 7.24 Sharp space solar cell conversion efficiency heritage. CONV., BSF, BSFR,
NRS/BSF (IBF) show conventional p-n junction, back-surface-field, back-surface-field-
reflector, non-reflective surface/back-surface field (improved back-surface field) structures
for Si space solar cells. AHES shows advanced high-efficiency solar cell structures for space
cells. BOL and EOL show beginning-of-life and end-of-life, respectively. All efficiencies
for 1 Sun AM 0 under standard conditions.
can be channelled into the lattice vibration mode which drives the defect motion:
EI = E A − ER .
41.1% efficiency under 454 Suns with In0.65 Ga0.35 P/In0.17 GaAs0.83 /Ge 3-J cells
(Bett et al., 2009), and Spectrolab reached 41.6% efficiency under 364 Suns with
lattice-matched InGaP/InGaAs/Ge 3-J cells (King et al., 2009). More recently,
Spire achieved 42.3% efficiency under 406 Suns with bifacial epitaxially grown
InGaP/GaAs/InGaAs three-junction cells (Wojtczuk et al., 2010) and Solar Junc-
tion (Solar Junction, 2011; Sabnis, 2012) reached 43.5% efficiency under 418 Suns
with lattice-matched InGaP/GaAs/GaInNAs 3-J cells. At the time of writing, the
world record is held by Sharp with their InGaP/GaAs/InGaAs 3-J solar cells, which
are 44.4% efficient at 302 Suns (Sharp, 2013; Sasaki, 2013; Yamaguchi and Luque,
2013; Green et al., 2013).
The 1 Sun efficiency of triple-junction solar cells is also improving. A world-
record efficiency (37.9%) at 1 Sun (AM 1.5 G) has recently been realised with
inverted epitaxially grown InGaP/GaAs/InGaAs three-junction cells by Sharp
(Sharp, 2013; Sasaki, 2013; Yamaguchi and Luque, 2013; Green et al., 2013).
Figure 7.25 shows the fabrication process of this cell (Takamoto et al., 2010;
Yoshida et al., 2011) and Fig. 7.26 shows its current–voltage curve. Figure 7.27
shows the chronological improvement in the conversion efficiencies of III–V com-
pound MJ solar cells under 1 Sun and concentrator conditions.
332 M. Yamaguchi
45
number of junctions III–V, concentrator
1-junction 44.4%
■
2-junction this time
40
3-junction
33.3%
efficiency / %
35.8%
30 previous
30.3% III–V, 1 Sun AM1.5G
JE
25
single Si, 1 Sun AM1.5G
20
1985 1990 1995 2000 2005 2010 2015
year
Figure 7.27 Chronological improvements in conversion efficiencies of III–V compound
multijunction solar cells under 1 Sun and concentrator conditions (Yamaguchi and Luque,
2013).
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch07 page 333
50
III−V multijunction
40
efficiency / %
30
thin-film Si
CIS
20 dye-sensitised
cryst. Si
organic
10
a-Si
0
1940 1960 1980 2000 2020 2040 2060
year
Figure 7.28 Predictions of future solar cell efficiencies (Goetzberger et al., 2001; Yam-
aguchi, 2004; updated by using Solar Cell Efficiency Tables [Green et al., 2013]).
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch07 page 334
334 M. Yamaguchi
1:Si crystal
2:thin-film
3:concentrator
100
2
Electricity Cost (Yen/kWh)
3
10
1
1995 2000 2005 2010 2015 2020 2025 2030
Year
Figure 7.29 Scenario of electricity cost reduction by developing concentrator solar cells
(Yamaguchi, 2003).
substrates and would have a theoretical efficiency of about 42% under 1 Sun AM
0, and over 47% under 500 Suns AM 1.5 G (Kurtz et al., 1997).
In conclusion, as shown in Fig. 7.29 (Yamaguchi, 2004), we would like to
contribute to commercialisation of CPV technologies as the third-generation PV
technologies in succession to the first-generation crystalline Si PV and the second-
generation thin-film PV technologies.
Acknowledgements
This work was partially supported by the Japanese New Energy and Industrial Tech-
nology Development Organization (NEDO) under METI (Ministry of Economy,
Trade and Industry). The author thanks members of the Toyota Institute, Sharp,
Daido Steel, JAXA, JAEA, the University of Tokyo, Meijo University, Kyushu
University and Miyazaki University for their collaboration and cooperation.
References
Ahrenkiel R. K., Keyes B. M., Durbin S. M. and Gray J. L. (1993), ‘Recombination lifetime
and performance of III–V compound photovoltaic devices’, Proc. 23rd IEEE Photo-
voltaic Specialists Conf., IEEE, New York, NY, pp. 42–51.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch07 page 335
Amano C., Sugiura H., Yamaguchi M. and Hane K. (1989), ‘Fabrication and numerical anal-
ysis of AlGaAs/GaAs tandem solar cells with tunnel interconnections’, IEEE Trans.
Electron Devices ED–36, 1026–1035.
Ando K., Amano C., Sugiura H., Yamaguchi M. and Saletes A. (1987), ‘Non-radiative
e-h recombination characteristics of mid-gap electron trap in Alx Ga1−x As (x = 0.4)
grown by molecular beam epitaxy’, Jpn. J. Appl. Phys. 26, L266–L269.
Araki K., Uozumi H., Egami T., Hiramatsu M., Miyazaki Y., Kemmoku Y., Akisawa A.,
Ekins-Daukes N. J., Lee H.-S. and Yamaguchi M. (2005), ‘Development of concentrator
modules with dome-shaped Fresnel lenses and triple-junction concentrator cells’,
Progr. Photovoltaics 13, 513–527.
Bertness K. A., Kurtz S. R., Friedman D. J., Kibbler A. E., Kramer C. and Olson J.
M. (1994), ‘29.5%-Efficiency GaInP/GaAs tandem solar cells’, Appl. Phys. Lett. 65,
989–991.
Bett A. W., Dimroth F., Guter W., Hoheisel R., Oliva E., Philips S. P., Schone J., Siefer G.,
Steiner M., Wekkeli A., Welser E., Meusel M., Kostler W. and Strobl G. (2009), ‘Highest
efficiency multijunction solar cell for terrestrial and space applications’, Proc. 24th
European Photovoltaic Solar Energy Conf., WIP, Munich, pp. 1–6.
Bowler D. L. and Wolf M. (1980), ‘Interactions of efficiency and material requirements for
terrestrial silicon solar cells’, IEEE T Compon. Hybr., CHMT–3, 464–472.
Brown M. R., Goldhammer L. J., Goodelle G. S., Lortz C. U., Perron J. N., Powe J. S.
and Schwartz J. A. (1997), ‘Characterization testing of dual junction GaInP2 /GaAs/Ge
solar cell assemblies’, Proc. 26th IEEE Photovoltaic Specialists Conf., IEEE, New
York, NY, pp. 805–810.
Carlin J. A., Hudait M. K., Ringel S. S., Wilt D. M., Clark E. B., Leitz C. W., Currie M.,
Langdo T. and Fitzgerald E. A. (2000), ‘High-efficiency GaAs-on-Si solar cells with
high Voc using graded GeSi bufers’, Proc. 28th IEEE Photovoltaic Specialists Conf.,
IEEE, New York, NY, pp. 1006–1011.
Chung B.-C., Virshup G. F., Hikido S. and Kaminar N. R. (1989), ‘27.6% efficiency (1 Sun,
air mass 1.5) monolithic Al0.37 Ga0.63 As/GaAs two-junction cascade solar cell with
prismatic cover glass’, Appl. Phys. Lett. 55, 1741–1743.
Fan J. C. C., Tsaur B. Y. and Palm B. J. (1982), ‘Optical design of high-efficiency tan-
dem cells’, Proc. 16th IEEE Photovoltaic Specialists Conf., IEEE, New York, NY,
pp. 692–701.
Flores C. (1983), ‘A three-terminal double junction GaAs/GaAlAs cascade solar cells’,
IEEE Electr. Device L. EDL–4, 96–99.
Fraas L. M., Avery J. E., Martin J., Sundaram V. S., Girard G., Dinh V. T., Davenport T. M.,
Yerkes J. W. and O’Neill M. J. (1990), ‘Over 35% efficient GaAs/GaSb tandem solar
cells’, IEEE Trans. Electr. Devices 37, 443–449.
Goetzberger G., Luther J. and Willeke G. (2002), ‘Solar cells: past, present, future’, Solar
Energy Mater. Solar Cells, 74, 1–11.
Green M., Emery K., Hishikawa Y., Warta W. and Dunlop E. D. (2013), ‘Solar efficiency
tables (Version 42)’, Progr. Photovoltaics 21, 827–837.
Hutchby J. A., Markunas R. J., Timmons M. L., Chiang P. K. and Bedair S. M. (1985), ‘A
review of multijunction concentrator solar cells’, Proc.18th IEEE Photovoltaic Spe-
cialists Conf., IEEE, New York, NY, pp. 20–27.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch07 page 336
336 M. Yamaguchi
Imaizumi M., Matsuda S., Kawakita S., Sumita T., Takamoto T., Ohshima T. and
Yamaguchi M. (2005), ‘Activity and current status of R&D on space solar cells in
Japan’, Progr. Photovoltaics 13, 529–543.
Jackson E. D. (1955), ‘Areas for improving of the semiconductor solar energy converter’,
Trans. Conf. on the Use of Solar Energy 5, University of Arizona Press, Tucson (1958),
pp. 122–126.
Khan A., Yamaguchi M., Bourgoin J. C. and Takamoto T. (2000), ‘Room-temperature
minority-carrier injection-enhanced recovery of radiation-induced defects in p-InGaP
and solar cells’, Appl. Phys. Lett. 76, 2559–2561.
King R. R., Boca A., Hong W., Liu X.-Q., Bhusari D., Larrabee D., Edmondson K. M.,
Law D. C., Fetzer C. M., Mesropian S. and Karam N. H. (2009), ‘Bandgap-engineered
architectures for high-efficiency multijunction concentrator solar cells’, Proc. 24th
European Photovoltaic Solar Energy Conf., WIP, Munich, pp. 55–61.
Kurtz S. R., Myers D. and Olson J. M. (1997), ‘Projected performance of three-and four-
junction device using GaAs and GaInP’, Proc. 26th IEEE Photovoltaic Specialists
Conf., IEEE, New York, NY, pp. 875–878.
Lamorte M. F. and Abbott D. H. (1980), ‘Computer modeling of a two-junction monolithic
cascade solar cell’, IEEE Trans. Electr. Devices ED–25, 831–840.
Lang D. V., Kimerling L. C. and Leung S. Y. (1976), ‘Recombination-enhanced annealing
of the E1 and E2 defect levels in 1-MeV-electron-irradiated n-GaAs’, J. Appl. Phys.
47, 3587–3591.
Letay G. and Bett A. W. (2001), ‘Theoretical investigation III–V multijunction solar cells’,
Proc. 17th European Photovoltaic Solar Energy Conf., WIP, Munich, pp. 178–181.
Loferski J. J. (1976), ‘Tandem photovoltaic solar cells and increased energy conversion
efficiency’, Conf. Record 12th IEEE Photovoltaic Specialists Conf., Baton Rouge.
IEEE Press, Piscataway, NJ, pp. 957–961.
Ludowise M. J., LaRue R. A., Borden P. G., Gregory P. E. and Dietz W. T. (1982), ‘High-
efficiency organometallic vapor phase epitaxy AlGaAs/GaAs monolithic cascade solar
cell using metal interconnects’, Appl. Phys. Lett. 41, 550–552.
Mitchell K. W. (1981), ‘High efficiency concentrator cells’, Conf. Record 15th IEEE
Photovoltaic Specialists Conf., Kissimmee. IEEE Press, Piscataway, NJ, pp. 142–146.
Nell M. E. and Barnett A. M. (1987), ‘The spectral p-n junction model for tandem solar-cell
design, IEEE Trans. Electr. Devices ED–34, 257–266.
Olson J. M., Kurtz S. R. and Kibbler A. E. (1990), ‘A 27.3% efficient Ga0.5 In0.5 P/GaAs
tandem solar cell’, Appl. Phys. Lett. 56, 623–625.
Sabnis V., Yuen H. and Wiemer M. (2012), ‘High-efficiency multijunction solar cells
employing dilute nitrides’, Proc. 9th Int. Conf. on Concentrating Photovoltaics Sys-
tems, Toledo, Spain.
Sasaki K., Agui T., Nakaido K., Takahashi N., Onitsuka R. and Takamoto T. (2013), ‘Devel-
opment of InGaP/GaAs/InGaAs inverted triple junction concentrator solar cells’, Proc.
9th Int. Conf. on Concentrating Photovoltaics Systems, Miyazaki, Japan.
Sharp (2013), http://www.sharp.co.jp.
Solar Junction (2011), http://www.sj-solar.com.
Sugiura H., Amano C., Yamamoto A. and Yamaguchi M. (1988), ‘Double heterostructure
GaAs tunnel junction for AlGaAs/GaAs tandem solar cells’, Jpn. J. Appl. Phys. 27,
269–272.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch07 page 337
Takamoto T., Agui T., Ikeda M. and Kurita E. (2000), ‘High-efficiency InGaP/GaAs tandem
solar cells on Ge substrates’, Proc. 28th IEEE Photovoltaic Specialists Conf., IEEE,
New York, NY, pp. 976–981.
Takamoto T., Agui T., Kamimura K., Kaneiwa M., Imaizumi M., Matsuda S. and
Yamaguchi M. (2003), ‘Multijunction solar cell technologies — high efficiency,
radiation resistance and concentrator applications’, Proc. 3rd World Conf. Photovoltaic
Energy Conversion, WCPEC-3, Osaka, pp. 581–586.
Takamoto T., Agui T., Yoshida A., Nakaido K., Juso H., Sasaki K., Nakamora K.,
Yamaguchi H., Kodama T., Washio H., Imaizumi M. and Takahashi M. (2010),
‘World’s highest efficiency triple-junction solar cells fabricated by inverted layers trans-
fer process’, Proc. 35th IEEE Photovoltaic Specialists Conf., IEEE, New York, NY,
pp. 412–417.
Takamoto T., Ikeda E., Kurita H. and Ohmori M. (1997a), ‘Over 30% efficient InGaP/GaAs
tandem solar cells’, Appl. Phys. Lett. 70, 381–383.
Takamoto T., Ikeda E., Kurita H., Ohmori M. and Yamaguchi M. (1997b), ‘Two-terminal
monolithic InGaP/GaAs tandem solar cells with a high conversion efficiency of over
30%’, Jpn. J. Appl. Phys. 36, 6215–6220.
Takamoto T., Kaneiwa M., Imaizumi M. and Yamaguchi M. (2005), ‘InGaP/GaAs-based
multijunction solar cells’, Progr. Photovoltaics 13, 495–511.
Takamoto T., Yamaguchi M., Ikeda E., Agui T., Kurita H. and Al-Jassim M. (1999), ‘Mecha-
nism of Zn and Si diffusion from highly doped tunnel junction for InGaP/GaAs tandem
solar cells’, J. Appl. Phys. 85, 1481–1485.
WBGU (German Advisory Council on Global Change) (2003), World in Transition —
Towards Sustainable Energy Systems, Earthscan, London, http://wbgu.de.
Wojtczuk S., Chiu P., Zhang X., Derkacs D., Harris C., Pulver D. and Timmons
M. (2010), ‘InGaP/GaAs/InGaAs 41% concentrator cells under bi-facial epi-
growth’, Proc. 35th IEEE Photovoltaic Specialists Conf., IEEE, New York, NY,
pp. 1259–1264.
Wolf M. (1960), ‘Limitations and possibilities for improvement of photovoltaic solar energy
converters’, Proc. Inst. Radio Engineers 48, 1246–1263.
Yamaguchi M. (2003), ‘III–V compound multijunction solar cells: present and future’, Solar
Energy Mater. Solar Cells, 75, 261–269.
Yamaguchi M. (2004), ‘Toward 50% efficiency III–V compound multijunction and concen-
trator solar cells’, Proc. 19th European Photovoltaic Solar Energy Conf., WIP, Munich,
xl–xlii.
Yamaguchi M. and Amano C. (1985), ‘Efficiency calculations of thin film GaAs solar cells
on Si substrates’, J. Appl. Phys. 58, 3601–3606.
Yamaguchi M., Amano C., Sugiura H. and Yamamoto A. (1987), ‘High-efficiency
AlGaAs/GaAs tandem solar cells’, Proc. 19th IEEE Photovoltaic Specialists Conf.,
IEEE, New York, NY, pp. 1484–1485.
Yamaguchi M. and Luque, A. (1999), ‘High efficiency and high concentration in photo-
voltaics’, IEEE Trans. Electron Devices 46, 2139–2144.
Yamaguchi M. and Luque A. (2013), ‘Outline of Europe–Japan collaborative research on
concentrator photovoltaics’, Proc. 39th IEEE Photovoltaic Specialists Conf., IEEE,
New York, NY.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch07 page 338
338 M. Yamaguchi
Yamaguchi M., Ohmachi Y., Oh’hara T., Kadota Y., Imaizumi M. and Matsuda S., (2001),
‘GaAs solar cells grown on Si substrates for space use’, Progr. Photovoltaics 9,
191–201.
Yamaguchi M., Okuda T., Taylor S. J., Takamoto T., Ikeda E. and Kurita H. (1997), ‘Superior
radiation-resistant properties of InGaP/GaAs tandem solar cells’, Appl. Phys. Lett.
70, 1566–1568.
Yamaguchi M., Warabisako T. and Sugiura H. (1994), ‘CBE as a breakthrough technology
for PV solar energy applications’, J. Crystal Growth 136, 29–36.
Yang M.-J., Yamaguchi M., Takamoto T., Ikeda E., Kurita H. and Ohmori M. (1997), ‘Photo-
luminescence analysis of InGaP top cells for high-efficiency multijunction solar cells’,
Solar Energy Mater. Solar Cells 45, 331–339.
Yoshida A., Agui T., Nakaido K., Murasawa K., Juso H., Sasaki K. and Takamoto T. (2011),
‘Development of InGaP/GaAs/InGaAs inverted triple junction solar cells for concen-
trator application’, Extended Abstracts of 21st International Photovoltaic Science and
Engineering Conf., Fukuoka, Japan, 28 November–2 December, 4B–4O–01.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch08 page 339
CHAPTER 8
ORGANIC PHOTOVOLTAICS
DAN CREDGINGTON
Cavendish Laboratory, JJ Thomson Avenue
Cambridge CB3 0HE, UK
[email protected]
8.1 Introduction
In this chapter, we move our focus away from the numerous elemental and crys-
talline semiconductors which dominate the existing technological landscape of
photovoltaics. For these materials, it is the specific arrangement of elements in a
(usually) crystalline lattice which gives rise to delocalised electronic states and
resulting band structure. The quality of the band structure therefore depends rather
sensitively on how such lattices are formed, and the necessary delocalisation of
electrons leads to materials that are highly sensitive to defects — including both
impurities and crystallite boundaries. The creation of high-performance materials
by this approach is an intrinsically intolerant, and therefore expensive, process. As
is detailed elsewhere in this volume, the primary brake on uptake of photovoltaics
for the last several decades has been the relatively high cost of silicon crystal, which
must be processed from a purified melt at extremely high temperature. While the
price of silicon crystal has reduced considerably as the demand for solar energy
has grown, routes that circumvent high-tolerance manufacturing may lead to even
cheaper photovoltaic devices.
An alternative approach is to utilise materials in which semiconducting
behaviour is intrinsic, rather than an emergent property of lattice formation.
Molecular semiconductors are such a class of materials, whereby pre-synthesised
molecules with desirable electronic structure are used to form the light-absorbing
layer of the photovoltaic device. Interactions between molecules lead to broaden-
ing of the molecular orbitals, but this is a relatively small perturbation, which does
not typically lead to three-dimensional extended electronic states. The advantage
339
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch08 page 340
340 D. Credgington
Figure 8.1 Schematics of the electronic orbitals of a) sp3 hybridisation; b) sp2 hybridis-
ation; and c) sp hybridisation, with examples showing the a) diamond, b) graphene and c)
heptayne, octayne and dodecayne linear acetylenic carbon structures, with various capping
groups.
which form the basis of the rich field of synthetic organic chemistry, examples of
which are shown in Fig. 8.2.
Hybridisation between the remaining 2 pz orbitals gives rise to weaker π
(bonding) or π ∗ (anti-bonding) orbitals, the former arising from a symmetric
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch08 page 342
342 D. Credgington
Figure 8.2 Chemical structures for a range of commonly used small-molecule and poly-
mer semiconductors. CuPc = Copper phthalocyanine; SubPc = boron subphthalocyanine
chloride.
Figure 8.3 Schematic of energy-level splitting and absorption in alkenes with increas-
ing conjugation length, highlighting the lowest-energy optical transitions from HOMO to
LUMO. The bandgap of polyacetylene depends on both the number of monomers n and the
effective conjugation length in the polymer. Arrows represent spin-paired electrons.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch08 page 344
344 D. Credgington
layers. Such ‘π −π stacking’ is important for transport (see Section 8.4.3) and also
leads to a lower optical bandgap. A combination of intra- and inter-molecular delo-
calisation can therefore lead to molecules with absorption in the visible spectral
range.
Since thermal excitation from π to π* is negligible for visible-light transitions,
such materials behave as intrinsic semiconductors with the ‘bandgap’ defined by
E g , the difference between the highest occupied molecular orbital (HOMO) and
lowest unoccupied molecular orbital (LUMO). Doping of the molecule, or injec-
tion of external charge, is therefore necessary to initiate and exploit conduction.
In addition, each conjugated region will not necessarily extend over the whole
molecule (particularly in the case of larger polymers), limiting the practical extent
of delocalisation. However, as we shall see in Section 8.4.4, numerous alternative
routes exist for engineering the HOMO–LUMO offset, and a key advantage of these
materials is that E g is a molecular property, so may be defined before the material
is incorporated into a photovoltaic device. In addition, most organic molecules
are direct-bandgap materials with low symmetry, and thus exhibit extremely high
absorption coefficients — a film of organic material 100–200 nm thick is usually
sufficient to achieve complete absorption at the primary π − π* transition. How-
ever, the molecular (rather than band-like) nature of absorption usually means that
the optical density is peaked rather than step-like at this transition — ideal for dyes,
but not for photovoltaic absorbers.
For example, pentacene ∼ 4, PPV ∼ 2 (Martens et al., 1999) whereas Si ∼ 12,
GaAs ∼ 13. In materials with high , dielectric screening reduces the effective
Coulomb attraction of the excited electron-hole pair, leading to very low binding
energies and characteristic radii greater than the material’s lattice spacing — i.e.
electrons and holes are delocalised over many unit cells. This subset of excitons,
described by and named after Gregory Wannier (Wannier, 1937), are characterised
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch08 page 345
by binding energies of a few kT at room temperature, and thus may undergo spon-
taneous dissociation into free charges via interactions with lattice phonons.
In materials with low , the coulomb attraction between the electron and hole is
not strongly screened, leading to tightly bound Frenkel-type excitons (named after
Jakov Frenkel [Frenkel, 1931]). These have binding energies of order 0.1–1 eV
and characteristic radii of the same order as the unit cell (for crystals) or individual
monomers/molecular units (for molecular materials). Dissociation of these excitons
into free charge carriers is not typically spontaneous at room temperature, and
must be achieved by other means, which we discuss in Section 8.3.1. If they do not
dissociate, excitons in organic materials typically exhibit luminescence lifetimes
of a few nanoseconds. Since one of the most obvious effects of encouraging exciton
dissociation is to quench this luminescence, the term ‘exciton quenching’ is often
used to describe processes leading to exciton dissociation, whether or not the
material in question is intrinsically luminescent.
• Organic materials are electronically ‘soft’
Charge carriers moving through an organic material are usually localised to a few
carbon atoms, and so act as large perturbations to the local electronic structure.
Polarisation of the local medium (enhanced by low ) and the breaking of local
conjugation by the occupancy of the π* anti-bonding state leads to geometric relax-
ation of the molecule. The fundamental charge carrier in organic materials is thus
a quasi-particle, comprising the charge self-localised by its associated molecular
distortion: this entity is called a polaron. The ‘binding’ energy of the polaron (as
compared with the energy of a free charge without distortion) is estimated to be of
order 100 meV, and is determined by the competition between the energy gained
by relaxation of the surrounding medium around the charged excitation and the
energy cost of decreasing delocalisation. A consequence of this self-localisation
is that the migration of polarons between polymer chains or conjugated regions
involves significant reorganisation energy, and is limited by the electronic coupling
between conjugated regions. This relaxation also produces a significant Stokes
shift between absorption and emission spectra, which is particularly beneficial for
reducing self-absorption in organic light-emitting diodes (LEDs), but represents
a potential loss of free energy in a photovoltaic. The extent of relaxation is gov-
erned by the physical properties of the particular molecule; thus physically ‘stiff’
materials tend to exhibit lower polaron binding energies and faster charge transport.
346 D. Credgington
additional thermal energy. This energy is associated with the geometric relaxation
of the ‘acceptor’ molecule combined with the geometric rebound of the ‘donor’
molecule. In addition, the probability for a carrier to transfer is also dependent on the
wave-function overlap between donor and acceptor states. Since organic molecules
(and particularly polymers) typically exhibit low symmetry, transport can be highly
anisotropic. For a stiff conjugated polymer, intra-chain transport will be faster than
transport in the π − π stacking direction, while lateral transport through saturated
side groups can be very poor. For example, intra-chain hole mobilities as high as
600 cm2 /Vs have been reported for ladder-type poly( p-phenylene) (Prins et al.,
2006), whereas measured mobilities are of order 1–10 cm2/Vs (Li et al., 2012)
in solid films, and depend strongly on polymer orientation (Lee et al., 2011).
Such transport also depends on temperature and electronic disorder, which we will
discuss in more detail in Section 8.4.2.
• Transport of excitons is disorder-limited
The migration of excitons through the semiconductor occurs via transfer of the
bound polaron pair from one conjugated molecule/chain segment to another
through either Förster or Dexter energy transfer. Förster transfer (Förster, 1948)
involves short-range dipole–dipole interactions equivalent to the exchange of a
virtual photon; Dexter transfer (Dexter, 1953) involves direct exchange of elec-
trons. Requiring only coupling between dipoles, Förster transfer of singlet excitons
is the dominant process in photovoltaics and is typically restricted to a range of
5–20 nm over the lifetime of the exciton (Halls et al., 1996; Stubinger and Brutting,
2001), depending on the material in question. This range is limited by the degree
of overlap between the material’s absorption and emission spectra (reduced by
polaronic relaxation), and by disorder in the energetic landscape accessible to the
exciton. Since Dexter transfer requires wave-function overlap between the donat-
ing and accepting species, it is usually dominant only in situations where optical
transitions are forbidden, for example in the migration of spin-triplet excitons.
are usually quite stiff due to the planarising effect of their π-bonding, solubilisa-
tion is usually achieved by attaching flexible saturated side groups along the main
backbone, whose chemistry determines the range of compatible solvents. For sim-
plicity, we will here refer to both the conjugated segments of a polymer chain, and
the conjugated carbon scaffold of organic small molecules as the ‘backbone’ of
the structure.
Many organic molecules may therefore be formed into inks and processed from
solution. This approach is of particular importance for depositing large semicon-
ducting molecules such as conjugated polymers, for which vacuum sublimation is
impossible. As solid films, these polymeric semiconductors are flexible and highly
resistant to fracture, lending the collection of technologies which exploit them the
umbrella term ‘plastic electronics’.
In solution, the freedom of organic molecules to explore a range of config-
urations is enhanced by the presence of the solvent. Deposition from solution is
therefore thermodynamically analogous to a rapid quench, as the evaporation of
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch08 page 348
348 D. Credgington
solvent molecules quickly restricts this freedom. Depending on the details of the
deposition (and particularly on the evaporation rate of the solvent), the resulting
material is typically left in a metastable configuration, with a mix of glassy and
more ordered regions. This effect is even more prevalent in systems comprising
mixtures of multiple materials. Since electronic structure is influenced by both
local conjugation length and intermolecular interactions, solution deposition leads
unavoidably to significant spatial and energetic disorder. This intrinsic disorder has
implications for many photovoltaic properties, as we discuss in Section 8.4.2.
Deposition of organic semiconductors from solution can take advantage of
the wide range of thin-film coating technologies already employed by existing
industries. These include inkjet and screen-printing for individual substrates and
extend to continuous roll-to-roll manufacture, for example using slot-die, gravure
printing or doctor-blading. The existence of mature, low-cost, large-area manu-
facturing routes means that scaling-up of organic semiconductor technology is
widely expected to be a relatively cheap process, if the materials set can also reach
maturity. We discuss the progress made to date in large-area OPV manufacture in
Section 8.5.6.
Figure 8.5 Conventional (a) and inverted (b) device structures, indicating stack order and
interlayer positions.
350 D. Credgington
also be used directly to split excitons, which was found to be consistent with the
electric-field-induced ionisation mechanism proposed by Onsager, and later refined
by Braun (Braun, 1984; Onsager, 1934). However, the primary limit on efficiency
remained the competition between the short exciton lifetime (a few ns) and the low
rate of spontaneous exciton dissociation under normal operating conditions.
Planar heterojunctions
A number of pieces of evidence pointed to a possible solution. In single-material
devices, dissociation was found to be more likely at surfaces or impurities, and in
the presence of water or oxygen. It was also discovered that electron transfer from
organic pigments to other semiconductors — such as from chlorophyll to ZnO —
could be efficient (Tributsch and Calvin, 1971). The implication of these obser-
vations was that exciton dissociation could be encouraged by allowing charges
to transfer to a nearby heterogeneous molecule. This was confirmed by Tang and
co-workers (Tang, 1986), who showed that the photocurrent of an organic photo-
voltaic could be increased by orders of magnitude if two different semiconducting
molecules were combined to form a bilayer, or ‘planar’ heterojunction (shown in
Fig. 8.7a). Rather than relying on thermal energy or an electric field to ionise the
excitons, this structure drives dissociation by using the gain in free energy associ-
ated with the transfer of an electron from the LUMO of the ‘donor’ semiconductor
to the LUMO of the ‘acceptor’ semiconductor (or the transfer of a hole from accep-
tor HOMO to donor HOMO). In the language of crystalline semiconductors, a type
II heterojunction is required. Figure 8.8 illustrates this process.
To a first approximation, the band-edge offset must be greater than or equal to
the exciton binding energy, a point we shall discuss in more detail in Section 8.4.4.
Figure 8.7 Heterojunction structures, showing schematics of: a) planar heterojunction; and
b) bulk heterojunction photoactive layers in a conventional structure.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch08 page 352
352 D. Credgington
By careful choice of materials — CuPc and the perylene derivative ‘PV’ in this case
(shown in Fig. 8.2) — devices with PCEs of around 1% were realised. In addition,
the use of two semiconductors opened the possibility of combining materials with
complementary absorption spectra in a single device.
Bulk heterojunctions
While a significant advance, the EQE of planar heterojunction devices peaked at
around 15%, since only those excitons able to diffuse to the heterojunction could
be dissociated. With the diffusion length of singlet excitons limited to approxi-
mately 10 nm, and absorption lengths of order 100 nm, we must often accept either
incomplete light harvesting or incomplete exciton harvesting when using such a
bilayer configuration.
This problem was solved in the mid-1990s with the introduction of the ‘bulk’
heterojunction (BHJ) structure (Halls et al., 1995; Yu et al., 1995), the philosophy
of which is to ‘bring the heterojunction to the exciton’. By blending the donor and
acceptor materials together in a single layer, it is possible to produce a distributed
heterojunction extending throughout the device, such that every newly generated
exciton is within one diffusion length of an interface (Fig. 8.7b).
The discovery of this approach coincided with the development of highly
effective soluble fullerene derivatives (small molecules based on buckyballs) as
acceptors (Sariciftci et al., 1992). These favoured the use of organic polymers
as donors, since the entangled polymer chains can prevent regions of the film
becoming electrically isolated, and can also encourage de-mixing (discussed in
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch08 page 353
354 D. Credgington
the structures of which are shown in Fig. 8.2. This means that the technology
has advanced from quantum yields of 10−3 to a point where nearly every exciton
can reach the heterointerface and dissociate, and under short-circuit conditions the
resulting carriers escape the device without loss.
8.3.2 Photovoltage
For any photovoltaic material, the absorption of photons with energy greater than
the semiconductor bandgap leads to thermalisation losses as electrons/holes relax to
the conduction/valence band edges (or LUMO/HOMO in this case). The maximum
free energy difference per carrier that remains after transport to the electrodes
defines the open-circuit voltage (Voc) of the cell. This is generally limited by either
the electrodes or the materials, as we now discuss.
Electrode-limited regime
After the development of the bulk heterojunction solved the problem of exciton
dissociation, several authors observed that a correlation existed between Voc and
the work function difference of the cell’s electrodes, (Brabec et al., 2002;
Mihailetchi et al., 2003; Mihailetchi et al., 2004). The implication was that the
energy difference between the donor HOMO and acceptor LUMO (often referred
to as the heterojunction bandgap, E HJ ) is reduced to when the carrier trans-
fers to the electrodes. This picture became known as the metal–insulator–metal
(MIM) model, and drove the development of materials to achieve increasingly
large . It is now common to use interlayers comprising soluble conducting
polymers (Aernouts et al., 2004), p-type oxides (Irwin et al., 2008; Zilberberg
et al., 2012) and surface treatments to increase the work function of the hole-
collecting electrode, and to use salts (Brabec et al., 2002), polymers (Zhou et al.,
2012), n-type oxides (Kim et al., 2006) and low work-function metals as inter-
layers to decrease the work function of the electron-collecting electrode. These
layers can provide an additional benefit if they act to increase the ‘selectivity’ of
the appropriate contact — i.e. to block the transport of electrons diffusing to the
hole-collecting contact, and vice versa.
Materials-limited regime
The drive to increase , and the development of lower-bandgap absorbers, has
revealed that where is greater than E HJ , it is the latter that limits the available
free energy (see Fig. 8.9a). Outside the MIM regime, applicable where < EHJ ,
numerous studies have shown correlations between the achievable Voc and E HJ , the
most well-known being the study by Scharber and co-workers (Scharber et al.,
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch08 page 355
2006) reproduced in Fig. 8.9b. The correlation is not exact — in all cases the
measured difference between donor HOMO and acceptor LUMO overestimates
the achievable Voc — but it provides an excellent design rule. In Section 8.4.5 we
discuss in more detail the origin of this discrepancy, and how the switch between
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch08 page 356
356 D. Credgington
UCLA-Sumitomo
12 Single Junction Mitsubishi Chemical
Tandem
Solarmer
Heliatek
10
Certified PCE (%)
Konarka Heliatek
8 Plextronics
University Groningen UCLA
Siemens Konarka
6
Sumitomo Chemical
4 Heliatek
NREL/Konarka
2 University Dresden
University Linz
0
2000 2004 2008 2012
Year
Figure 8.10 Progress of certified organic photovoltaic efficiencies over time since 2000,
including both single-junction and tandem devices. Labels indicate institution/company of
origin. Data from Green et al. (2013).
Bilayer heterojunctions
Bilayers are mainly of interest for studying interface physics in a well-defined
system, meaning that thermally evaporated small molecules are particularly well
suited to this architecture. Solution-processed bilayers require either strict solvent
orthogonality (i.e. two materials with mutually exclusive solubility) or a means
of rendering one layer insoluble before deposition of the next — for instance by
cross-linking (Png et al., 2010) or removal of solubilising groups (Bradley, 1987).
Failure to do so risks rapid mixing of the bilayer components. Even where this
is initially prevented, inter-diffusion of materials over time is often unavoidable
(Treat et al., 2011).
The best small-molecule absorbers remain the phthalocyanine and sub-
phthalocyanine derivatives (see Fig. 8.2), which achieve good light harvesting
in exceptionally thin (∼10–15 nm) films. When combined with the acceptor C60 ,
PCEs of order 3–4% are achievable (Lin et al., 2012; Peumans and Forrest, 2001).
358 D. Credgington
Commercial developers
For the last two decades, research into new materials and device optimisation has
been undertaken mainly within university research laboratories. It is worth noting
that within the last few years, this trend has changed: the most efficient organic
photovoltaics now reported are produced almost exclusively by commercial devel-
opers. As noted in connection with Fig. 8.10, the record for any OPV device cur-
rently lies with Heliatek (2013) and their 12%-efficient evaporated small-molecule
tandem solar cell.
Mitsubishi Chemical have also achieved certified PCEs of 11% with a
solution-processed small-molecule heterojunction, which is reported to be a
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch08 page 359
single-junction device (Green et al., 2013). However, in all cases the materials
and device structure are not known. It is therefore not clear in which direction the
future of the OPV field lies, but it is certain that the materials on which it is based
are still far from maturity.
G = H − TS (8.1)
360 D. Credgington
where G X is the free energy associated with phase X, G local captures the changes
to the local chemistry and motions of the mixed units, including any changes
to the local entropy due to volume changes, and Strans captures the change in
translational entropy of the system. By assuming that G local arises primarily from
enthalpic interactions between a given molecule and the mean field originating from
the rest of the mixture, it may be shown that
φA φB
G mix = RT V χφA φB + ln φA + ln φB (8.3)
NA NB
where V is the volume of the system, R the molar gas constant, φA and φB the
volume fractions of component A and B, and NA and NB their degrees of poly-
merisation. χ is the dimensionless ‘Flory–Huggins’ parameter, which represents
the change in local free energy per reference unit (monomer, in this case). The
first term represents local enthalpy change, while the second and third terms repre-
sent entropy change and will be negative. Homogeneous mixing will occur only if
G mix < 0, so compatible molecules (χ < 0) will always mix, but incompatible
molecules (χ > 0) will only do so for sufficiently small N. This implies that a
homogeneous mix of dissimilar polymers can usually lower its free energy via
spinodal decomposition into a phase-separated morphology, with little entropic
penalty. Figure 8.11a shows an example of such a morphology.
Real blend systems are always more complicated, as illustrated in Fig. 8.11c,
in particular because it has been observed empirically that at least one semicrys-
talline component is required for a device to be efficient. If phase separation can
also be driven by crystallisation, then even small molecules may spontaneously
de-mix, given sufficient thermodynamic freedom. For example, the small size of
fullerene acceptors means they are usually miscible in amorphous conjugated poly-
mers (or within the amorphous fraction of semicrystalline polymers) up to around
50% by volume, as might be expected from a simple Flory–Huggins description.
Beyond this miscibility threshold, purer fullerene domains develop which may
subsequently crystallise, adding significant complexity to the system.
For a two-component semicrystalline blend at least five phases are therefore
expected: two pure phases per component, one crystalline, one amorphous, plus a
mixed amorphous phase. The widely studied P3HT:PCBM system exhibits such
behaviour, with the final phase structure and domain size depending crucially on
the thermodynamic trajectory. The somewhat unusual system of poly(2,5-bis(3-
hexadecylthiophen-2-yl)thieno[3,2-b]thiophene) (PBTTT):PCBM adds a sixth
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch08 page 361
Figure 8.11 a) Topography (vertical range 81 nm) of a (1:1) PFO/F8BT thin film; b) F8BT
fluorescence (arbitrary scale) from the same region of the blend, illustrating the classic
phase-separated structure of a de-mixed polymer:polymer blend. Adapted from Chappell
et al. (2003). c) Sketch illustrating the much finer microstructure formed by the blend of
semicrystalline P3HT (lines) and PCBM (dots) with increasing fullerene loading: (i) with
no PCBM; (ii) with PCBM dissolved in the amorphous fraction of the P3HT film without
disrupting the typical crystallite spacing (L P , of order 10 nm); (iii) with sufficient PCBM
inclusion to swell the amorphous phase and increase L P ; (iv) saturation of amorphous-
phase swelling and the onset of PCBM aggregation between P3HT crystallites. Adapted
from Kohn et al. (2013).
362 D. Credgington
Figure 8.12 a) Phase diagram of the pBTTT/PCBM/co-crystal system, showing two eutec-
tics: at ∼1 wt% PCBM (pBTTT/co-crystal binary) and 43 wt% PCBM (co-crystal/PCBM
binary; b Photoluminescence quenching (PLQ) is correlated with the onset of co-crystal for-
mation beyond the first eutectic, as indicated by Ht(co−crystal), whereas long-lived charge
generation (OD) and collection (JSC) are correlated with the formation of crystalline
PCBM domains beyond the second eutectic. Reprinted from Jamieson et al. (2012).
Top-down patterning
The difficulty of generating a beneficial heterojunction structure using thermo-
dynamics alone has led to interest in enforcing a particular nanostructure on the
blend. To achieve this, a huge variety of techniques have been explored for pat-
terning organic semiconductors (Brédas et al., 2009). These include:
364 D. Credgington
Criteria 1 and 2 above improve the selectivity of the contacts, and reduce the
need for interlayer materials to be carrier-blocking. Criteria 3 and 5 are particularly
difficult to achieve simultaneously, since even in semicrystalline materials excitons
rarely diffuse much further than 10 nm, while criterion 6 requires a film thickness
of ∼200 nm. In principle, these requirements dictate an interdigitated structure of
pure donor and acceptor material with a very high aspect ratio, which has been
attempted by a variety of means using the techniques outlined above.
However, recent improvements in polymer:fullerene solar cell performance
have shown that neither pure donor and acceptor phases, nor well-structured het-
erojunction networks, are necessary for efficient operation. Regions of intermixed
donor and acceptor relax the need for single-phase percolation pathways, and sep-
arate excitons very efficiently. There is also evidence, discussed in Sections 8.4.4
and 8.4.5, that a hierarchical structure may provide an energetic driving force to
move carriers away from the interface. The assumption that a pure two-component
bulk heterojunction is ideal is therefore likely to be too simplistic, and the inclusion
of mixed or amorphous phases may be necessary for efficient operation.
Configurational disorder
Although all small organic molecules, or monomers, of a given material are struc-
turally identical, electronic interactions between two neighbouring molecules are
strongly dependent on their relative position (Linares et al., 2010). Since as-
deposited films are usually amorphous, or at most semicrystalline, the relative
position and orientation of neighbouring molecules varies significantly through-
out the film. This spatial disorder leads to highly anisotropic electronic coupling
between molecules, which impacts on charge transport. Similarly, since coupling
strengths between molecules affect the energetics of intermolecular and delocalised
states, this spatial disorder may contribute to energetic disorder, as each molecule
experiences a slightly different local environment.
Conformational disorder
For flexible molecules, variation in conformation also leads to spatial and energetic
disorder. This is particularly applicable to semiconducting polymers, where charge
delocalisation is dependent on conjugation length. Kinks or twists in the polymer
chain will break conjugation, limiting the extent of delocalisation and narrowing
the molecular ‘bands’. Chain-extended polymers are therefore optically distinct
from coiled polymers within the same sample (Clark et al., 2009), meaning that
variations in processing history and polydispersity, as well as random variation in
conjugation length, result in additional energetic disorder.
Dynamic disorder
The mechanisms above describe ‘static’ disorder, which will be valid at low
temperatures where molecular positions are fixed. At elevated temperature,
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch08 page 366
366 D. Credgington
Figure 8.13 Schematic illustration of charge (circles) hopping between localised states
(lines). Hopping from site A to site B represents nearest-neighbour hopping, while hopping
to sites C or D represents variable-range hopping, with lower energetic activation competing
with lower tunnelling probability.
368 D. Credgington
Figure 8.14 Models for transport in a disordered (electron) density of states. Localised
states are indicated in red, delocalised states in grey. a) Fully localised Gaussian DoS;
b) Band-like DoS with localised donor-like states within the bandgap; c) Transition from
localised to delocalised states within a continuous DoS, demarcated by an abrupt mobility
edge.
transition probabilities will be filled first, implying that the mobility of subsequent
carriers progressively increases. Such a modification was developed by Pasveer
and co-workers (Pasveer et al., 2005).
370 D. Credgington
order 107 V m−1 . Carrier drift is therefore a significant process and usually domi-
nates even at modest applied voltage, such that forward-biased organic diodes are
usually found to be space-charge limited (Mihailetchi et al., 2005).
Therefore, complete drift–diffusion simulations are required to capture the
main features of device operation in the power-generating quadrant, solving (in
the 1D case)
dn(x)
i n = i diffusion + i drift = q D + qµn E(x)n(x) (8.9)
dx
under the continuity condition
di n (x)
+ G(x) − R(x) = 0 (8.10)
dx
where i n is the current density due to electrons, µn is the electron mobility, G
and R are local generation and recombination rates for electrons, and we assume
equivalent expressions for holes with a requirement for charge neutrality. It is usual
to base such a model on an ‘effective medium’ representing the combined contri-
bution of each blend component, which can be made more realistic by including
localised variations in the DoS or separate contributions from different thermody-
namic phases.
In practice, obtaining sufficiently detailed data over a range of time and temper-
ature scales to isolate all of the internal processes contributing to charge transport
is extremely difficult (Bässler and Kohler, 2012).
Figure 8.15 AM1.5G solar spectrum and associated normalised cumulative photocurrent,
assuming that all photons below the given wavelength are converted to carriers and collected.
The absorption spectra of P3HT and PCPDTBT thin films are included for comparison.
systems, changing the acceptor is more difficult. Since C60 and its soluble deriva-
tives have spherical symmetry, their extinction coefficients are very low in the
visible range. Substituting C60 by the prolate C70 fullerene significantly increases
absorption in the visible and ultraviolet (UV), but at higher materials cost. To
date, it has not been possible to replace fullerenes successfully with more highly
absorbing acceptor materials.
The primary responsibility for absorption thus lies with the donor material.
Increasing conjugation lengths to reduce the HOMO–LUMO gap can only be
taken so far before molecules become too large or too stiff to deposit. Equally,
including heteroatoms within the (primarily hydrocarbon) organic backbone shifts
both HOMO and LUMO energies, but not necessarily their offset (Son et al.,
2011). This problem has been solved by coupling together electron-rich and
electron-withdrawing units in a ‘push-pull’ (or, confusingly ‘donor-acceptor’) con-
figuration, which can lead to the molecular HOMO states resting primarily on
the electron-rich unit, while the molecular LUMO states reside primarily on the
electron-withdrawing unit.
Figure 8.16 shows an example of this behaviour, illustrating the shift in electron
density for a 3CDTBT oligomer (analogous to the backbone of PCDTBT) from the
cyclopentadithiophene ‘push’ units to the benzothiadiazole ‘pull’ units. The result
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch08 page 372
372 D. Credgington
Figure 8.16 Shift in electron density on a 3CDTBT oligomer on excitation from the
ground (HOMO) to the first excited state (LUMO), calculated using time-dependent den-
sity functional theory. Electron density moves from red to blue regions. Image courtesy of
Sheridan Few.
has been a class of push-pull small molecules and polymers with absorption now
extending to beyond 900 nm (Bronstein et al., 2011; Mühlbacher et al., 2006; Zhong
et al., 2013). An additional benefit is that the higher-lying electronic transitions are
still accessible, which can broaden the absorption profile of such materials over
much of the visible range. The best performing polymer:fullerene cells utilise the
push-pull polymer PTB7 in combination with PC70BM to achieve >70% EQE from
400 to 750 nm, resulting in short-circuit current densities in excess of 17 mA cm−2
(He et al., 2012).
Exciton dissociation
While absorption of photons is a necessary precursor to current generation, the early
work on single-material devices showed that it was not sufficient. The necessity
for a type 2 (staggered) heterojunction indicates that a driving force is required to
separate charge, which represents a loss of available free energy. For example, in
the P3HT:PCBM system, around 800 meV of free energy is lost per carrier during
dissociation. Considerable effort has been expended to understand how this energy
loss may be minimised, while maintaining good separation efficiency (Clarke and
Durrant, 2010). We illustrate the main processes in Fig. 8.17, and refer to the rates
for those processes below.
When excitons reach the heterointerface, charge transfer from donor to accep-
tor is exceptionally fast, with the process complete within 50 fs — i.e. k CT kPL .
Studies of blend photoluminescence show that only 5–10 wt% of a good acceptor
mixed homogeneously with a donor is necessary to quench over 90% of exci-
tons. However, the result of such quenching is not necessarily dissociated charges.
The low dielectric constant of organic semiconductors means that the exciton may
remain bound across the heterointerface in a charge-transfer (CT) state. The spatial
separation necessitated by the interface means the binding energy of the CT state
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch08 page 373
Figure 8.17 Energy-level diagram illustrating the competing processes (with rates k i )
involved in photogeneration following absorption of a photon with energy hν. Exciton
diffusion and intersystem crossing of the donor singlet are ignored. Solid arrows repre-
sent processes increasing charge separation; dashed arrows represent relaxation towards the
ground state S0 . kCT : Dissociation of relaxed (or hot, kCT∗ ) singlet excitons (S) to form
hot interfacial charge-transfer (CT) states. kCS : Dissociation of relaxed (or hot, kCS∗ and
kCS∗∗ ) CT states to form fully dissociated charge-separated (CS) states. kISC : Intersystem
crossing between singlet and triplet CT states. k triplet : Generation of molecular triplet exci-
tons from triplet CT states, including possible re-dissociation at the heterojunction. k PL/Ph :
Relaxation of singlet/triplet excitons to the ground state. k GR : Geminate recombination
of CT states. kNGR : Non-geminate recombination of dissociated carriers, either directly
to the ground state or, more likely, via re-formation of CT states at the heterointerface.
k therm : Thermal relaxation to the corresponding lowest energy excited state. Since the CT
state involves partially separated carriers, there is unlikely to be a significant difference in
exchange interaction between 1 CT and 3 CT. Adapted from Clarke and Durrant (2010).
will be less than that of the molecular exciton, but still greater than the available
thermal energy; estimates for CT binding energies vary from around 0.1 to 0.5 eV
depending on the assumed separation and local polarisation. Excitons occupying
singlet CT states are observed to decay to the ground state within a few nanosec-
onds (a process usually termed geminate recombination, kGR in Fig. 8.17). While
the majority of CT states decay non-radiatively, some couple radiatively to the
ground state and produce CT luminescence, which provides a useful tool for their
study (Tvingstedt et al., 2009; Vandewal et al., 2009). The mechanism by which
CT excitons avoid geminate recombination and dissociate remains unclear, but
several routes have been proposed, as we now discuss.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch08 page 374
374 D. Credgington
Field-assisted ionisation
In many organic systems, macroscopic electric fields can distort the local energy
landscape and enhance separation efficiency. This process, observed in single-
component systems (Dick et al., 1994; Silinsh et al., 1974), polymer:polymer
systems (Gonzalez-Rabade et al., 2009), and some systems with fullerene acceptors
(Credgington et al., 2012; Veldman et al., 2008), is usually understood in terms
of Onsager–Braun theory. This describes the separation of electrons and holes in a
dielectric medium by calculating the probability that a Brownian random walk will
separate an electron-hole pair by more than their Coulomb capture radius — defined
by Onsager as the distance at which the Coulomb binding energy equals the thermal
energy kT. Excess energy provided by the incident photon (or heterojunction) would
generate an initially ‘hot’ carrier, and thus increase the chance of exceeding this
radius. Any macroscopic electric field will also lower the Coulomb potential in
the direction of the field, and so aid dissociation. Within this theory, the rate of
dissociation for a geminate pair kCS as a function of electric field E was derived
by Braun (1984) as:
3 µ
q b2 b3
kCS (E) = exp(−E/kT ) 1 + b + + + · · · (8.11)
4π s
a 3 3 18
where µ
is the spatially averaged sum of the electron and hole mobilities, s
376 D. Credgington
the chlorophyll absorbers. Each step costs energy, but decreases the associated
back-recombination rate. Considerable effort has therefore been directed towards
including spacer molecules or groups at the heterointerface to increase the electron-
hole separation of the CT state and thus increase its lifetime. This necessarily
reduces the rate of CT state formation, but since there are no competing processes
on the timescales involved, this need not be restrictive.
A related approach may already have been implemented empirically in hier-
archical polymer:fullerene bulk heterojunctions, where amorphous and crystalline
regions exist together. Since crystalline phases typically exhibit narrower bandgaps
than more amorphous phases, the existence of a crystalline region adjacent to the
more amorphous interface region may also drive carriers away from the interface
and thus reduce the geminate recombination rate (Jamieson et al., 2012).
378 D. Credgington
Metal–insulator–metal regime
In the metal–insulator–metal (MIM) regime, the electrode work functions lie within
the heterojunction bandgap. Since organic semiconductors are usually intrinsic,
equilibration between the electrodes is likely to deplete the entire semiconduc-
tor thickness, preserving the mismatch between the heterojunction energy levels
and the electrode work functions. Carriers reaching an electrode must therefore
thermalise to the work function, reducing their enthalpy. As such, the quasi-Fermi
energies in the bulk are limited by , giving
1
Voc = (h − e ) − V (8.12)
q
where e/ h are the work functions of the electron/hole extracting contacts,
respectively.
Materials-limited regime
If the electrode work functions lie outside the heterojunction bandgap, charge may
be transferred from the electrodes to empty states in the semiconductor, caus-
ing sharp band bending. This allows equilibration between the electrodes and the
HOMO/LUMO levels of the heterojunction without the build-up of space charge
in the majority of the semiconductor (provided the offset is not too great). In this
regime, the quasi-Fermi energies at the contacts are ‘pinned’ to states in the semi-
conductor (Crispin et al., 2006). For well-defined (donor) HOMO and (acceptor)
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch08 page 379
LUMO levels, Voc will be primarily governed by the relative difference between
these states, which we have termed E HJ . These energies are usually estimated
from pure donor or acceptor films using electrochemical or spectroscopic tech-
niques such as cyclic voltammetry or ultraviolet photoelectron spectroscopy. As
such, it is common for the measured quantities Ionisation Potential (IP) and Elec-
tron Affinity (EA) to be quoted instead of E HOMO and E LUMO . The latter are not
generally known, since blending donor and acceptor together may induce a change
in energetics, either through changes in molecular packing or the introduction of
interfacial dipoles. Where these changes may be ignored, however, we expect
1
Voc = (IPdonor − EAacceptor ) − V (8.13)
q
which is a widely observed correlation. A more accurate measure of the HOMO–
LUMO gap may be obtained by using the weak luminescence of the CT state as
an in situ probe of E HJ , thereby implicitly including the effects of blending. The
CT energy (E CT ) may be estimated by measuring either the electroluminescence
of the blend (which will be dominated by CT emission), or by isolating sub-
bandgap absorption features arising from CT states. The most accurate correlation
between energetics and open-circuit voltage is therefore given by (Faist et al., 2012;
Vandewal et al., 2008)
1
Voc = E CT −V (8.14)
q
where V in equation (8.14) is reduced by any contribution from the coulomb
binding energy of the CT state.
380 D. Credgington
Figure 8.18 Schematic illustrating the effect of DoS width on achievable Voc . For a fixed
number of photoexcited carriers, the wider LUMO DoS (a) exhibits a lower electron quasi-
Fermi energy E F n1 than the narrower DoS (b) with quasi-Fermi energy E F n2 . Both LUMO
distributions share a common centre.
loss will dominate V (Blakesley and Neher, 2011). At constant carrier density,
V will therefore be determined by the extent of disorder, as illustrated in Fig. 8.18.
The carrier density in a solar cell under operation is not a constant, however. At
open circuit, where no net current flows out of the device, the internal carrier density
will be set by the equilibrium between carrier generation and carrier recombina-
tion. Within a real solar cell, numerous additional recombination pathways exist
beyond the radiative recombination described in the Shockley–Queisser approach.
These include geminate recombination of CT excitons, collection of carriers at
the ‘wrong’ electrode, non-radiative recombination between dissociated polarons
(non-geminate recombination), recombination through trap sites and direct leak-
age currents, all of which can be important in organic solar cells and all of which
can vary in severity between devices (Credgington and Durrant, 2012).
A general expression for the open-circuit voltage must therefore include not
only the average heterojunction bandgap, but also the characteristic disorder width
and the severity of recombination losses. For a device limited by both geminate
and Langevin-type recombination (see below) and with fixed DoS N (assumed to
be the same for electrons and holes), Koster et al. (2005) derived the expression
E HJ kT (1 − P)γ N 2
Voc = − ln (8.15)
q q PG
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch08 page 381
where G is the generation rate of excitons per unit volume, P their dissociation
probability and γ a measure of the likelihood of Langevin recombination. Includ-
ing the effects of a general, disordered DoS and more complex recombination
mechanisms is not straightforward (Kirchartz et al., 2011), but in the case where
the dominant recombination process gives rise to a rate which varies as a power law
in carrier density, and the DoS varies exponentially with energy, the open-circuit
voltage may be expressed as
E HJ ηkT JBI
Voc = − ln (8.16)
q q Jgen
where η is an experimentally determined function of the DoS shape and recombina-
tion order, and JBI is related to the recombination rate at V = E HJ /q (Credgington
and Durrant, 2012).
382 D. Credgington
384 D. Credgington
uniquely sets the efficiency of an OPV at its maximum power point. This is deter-
mined by how EQE varies with internal field — i.e. the fill factor of the cell.
As we have outlined, the losses within any heterojunction solar cell arise from
either: (i) Relaxation of excitons which fail to diffuse to, and separate at, a het-
erojunction; (ii) Recombination of geminate pairs formed at the heterojunction
which fail to fully dissociate; or (iii) Recombination of fully dissociated carri-
ers via surface recombination, Langevin-type bulk recombination and SRH-type
recombination.
386 D. Credgington
8.5.1 Stability
Device lifetime is a critical concern for any commercial photovoltaic technology.
With the development of lab-scale OPVs with PCEs exceeding 10%, significant
attention has been focussed on the stability of organic devices. To compete directly
with existing technology, lifetimes approaching 100,000 hours under illumination
are sought. As with device performance, the rate of progress in this area has been
rapid, meaning that the state-of-the art described here will quickly become out-
of-date. We therefore outline the recent progress in developing stable devices,
and refer the reader to the series of reviews by Krebs and co-workers for a more
comprehensive outlook (Jørgensen et al., 2012). In particular, we do not consider
the interplay between stability and manufacturing process and the evolution of
encapsulation methods, which are not unique to organic photovoltaics.
In the following, we divide our discussion by degradation target, beginning
with the photoactive layer and extending to the electrodes, including the impact of
interfacial layers.
388 D. Credgington
Figure 8.20 ‘Rule-of-thumb’ photostability ranking of donor and acceptor monomers for
OPV polymers. Adapted from Manceau et al. (2011).
390 D. Credgington
achieve the good efficiency, annealing has been shown to be detrimental for many
better performing systems. It therefore remains to be seen whether thermo-cleaving
can be achieved at temperatures low enough to avoid this process.
An advantage of all of the approaches outlined above is that they also render
the blend insoluble, allowing for multiple thin-film stacks to be deposited without
the need for orthogonal solvents. This ability is crucial for developing multijunction
cells, as described in Section 8.5.2.
Electron-collecting electrode
The photoactive layer is not the only device component susceptible to degrada-
tion. Historically, the electron-collecting electrode has been formed of an evap-
orated metal. The requirement for this metal to have a low work function (see
Section 8.4.5) means that highly reactive materials such as calcium and aluminium
are needed. Such metals react in the presence of water or oxygen, creating an insu-
lating metal oxide layer which grows from pinholes and the edges of the electrode,
eventually spreading over the entire device. As the oxide increases in thickness,
the transport barrier created eventually impedes the extraction of electrons from
the entire active layer (Glatthaar et al., 2007).
One method to circumvent this problem is to use a less reactive, high-work-
function metal such as Ag or Au. These must be combined with a second mate-
rial, such as a conducting metal oxide, to achieve a low work function at the
organic:electrode interface. An alternative is to use an inverted structure, such that
Ag or Au, sometimes combined with a high-work-function interlayer, acts as the
hole-collecting electrode while a stable conductive oxide such as TiO2 is used to
extract electrons. The use of oxides as low-work-function electrodes has driven
interest in inverted device structures, since they are intrinsically stable to oxygen
degradation. In addition, while somewhat less conductive than TiO2 , solution-
processed titanium sub-oxide (TiOx ) has been shown to act as an oxygen ‘getter’
which also improves the photostability of the active layer (Li et al., 2011).
Hole-collecting electrode
While interfaces between ITO and organic materials have been shown to impact
stability, its relatively low work function means that it is unusual to use ITO without
a high-work-function hole-transporting layer — PEDOT:PSS (Fig. 8.6) being by
far the most widely used. However, PEDOT:PSS has been implicated in device
degradation through several mechanisms.
Since the conductivity of PEDOT:PSS relies on p-doping of the PEDOT
chain by the charged PSS component, phase separation, structural degradation
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch08 page 391
and electrochemical reactions that interfere with this charge exchange reduce the
conductivity of the PEDOT:PSS layer over time (Vitoratos et al., 2009). Commer-
cial PEDOT:PSS formulations can be highly acidic, such that ambient humidity can
significantly increase the rate of corrosion of other material components — partic-
ularly low-work-function metals, but also ITO (de Jong et al., 2000). PEDOT:PSS
is also highly hygroscopic, enhancing this effect even when minimal atmospheric
water is present.
As a result, there have been significant efforts to try and replace PEDOT:PSS
with a stable, transparent, high-work-function hole conductor. Examples of possi-
ble replacements include vacuum-deposited NiO (Irwin et al., 2008), and solution-
or vacuum-deposited MoO3 (Zilberberg et al., 2012). Both have been shown to
significantly enhance the stability of devices, particularly under humid conditions.
As in the previous section, an alternative approach is to switch to an inverted
structure such that a transparent electron-collecting electrode is used, and holes are
collected through a stable, high-work-function metal (Ag or Au, potentially with
an additional interlayer).
A huge number of other electron and hole-transporting interlayers have been
examined in the OPV and OLED fields, though many have not been assessed for
stability. For details, we refer the reader to the existing literature (Steim et al.,
2010).
392 D. Credgington
394 D. Credgington
396 D. Credgington
Figure 8.22 Singlet-fission processes and triplet harvesting. a) Schematic of indirect (blue)
and direct (green) pathways for the fission of one singlet into two triplets, which are arranged
on neighbouring chromophores in a pure singlet configuration. Adapted from Smith and
Michl (2010). b) Photogeneration in a singlet fission device: i) singlet exciton formation
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch08 page 397
Figure 8.22 (continued from caption on previous page) in the donor following absorption
of a Singlet-fission processes and triplet harvesting. a) Schematic of indirect (blue) and
direct (green) pathways for the fission of one singlet into two triplets, which are arranged on
neighbouring chromophores in a pure singlet configuration. Adapted from Smith and Michl
(2010). b) Photogeneration in a singlet fission device: i) singlet exciton formation in the
donor following absorption of a photon with energy greater than the donor bandgap E D ; ii)
singlet fission resulting in two coupled triplet excitons; iii) diffusion of triplet excitons to the
heterojunction forming two triplet CT states, which subsequently dissociate; iv) collection of
dissociated charges arising from the dissociated triplets and absorption of low-energy pho-
tons in the acceptor. Free-carrier generation is typically spontaneous in inorganic acceptors.
Solid lines represent single-carrier molecular energy levels. Dotted lines represent excitonic
‘virtual’ energy levels where the exciton binding energy has been schematically included.
Dashed lines indicate coulombically bound carriers. Interlayers and competing processes
have been ignored for clarity.
8.5.6 Scale-up
The increasing activity of industrial research groups in the OPV field is paral-
leled by increasing focus on applying scalable thin-film coating techniques to their
manufacture, and developing design rules for taking lab-scale devices to large
scale production. While much of this activity occurs within the commercial sector,
meaning that limited information is publicly available, scaling up OPV technology
requires several key transitions. We briefly outline these here, and refer the reader
to the growing literature on OPV scale-up for further details (Krebs, 2009; Krebs
et al., 2009).
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch08 page 398
398 D. Credgington
Scalable deposition
The first such transition is to move away from small-area serial production tech-
niques such as spin-casting. While suitable for lab-scale devices, this requires rela-
tively high solution wastage. In addition, variations in device properties result from
changes in evaporation rate, wet-film thickness, shear rates and nucleation surfaces
for crystallisation, which in turn depend on the spin rate, wetting behaviour and
solution viscosity. Spin-casting can be very reproducible within a given laboratory
environment for a given material batch, but it is very difficult to know the exact
parameters of a deposition such that it can be reproduced in a different location.
Large-area, roll-to-roll (R2R) compatible techniques are therefore preferred
for scalable and reproducible manufacture, as illustrated in Fig. 8.23. Numerous
processes have been explored, including screen-printing, doctor blading, wire-bar
coating, inkjet printing, knife-over-edge coating, slot-die coating, gravure printing
and spray coating (Krebs, 2009). Roll-to-roll is particularly appealing for solution-
based techniques since in-line assays can be used to correct for variations in film
formation caused by changing ambient conditions and solution composition.
In terms of reproducibility, it may be that the advantage lies with all-vacuum
R2R processing of thermally evaporated small molecules, as has been developed
by Heliatek (2013). While thermal evaporation under vacuum is a more technically
intensive route than solution processing, provided that all processing steps can be
Figure 8.23 A roll of printed OPV modules produced in a continuous R2R process. The
inset shows the layer stack, comprising PET-ITO-ZnO-P3HT:PCBM-PEDOT:PSS-printed
silver paint and encapsulated using R2R lamination post-production. Reproduced from
Krebs et al. (2010).
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch08 page 399
Large-area devices
The primary limit on efficient scale-up of OPV module size is series resistance.
Since organic semiconductors are low-mobility materials, it is not currently fea-
sible to employ laterally spaced high-conductivity electrodes, which are used in
crystalline solar cells to collect carriers while allowing the passage of light. It is
thus necessary to employ at least one transparent electrode with complete coverage
of the photoactive layer. The only material yet found to combine high optical trans-
mission (>80%) with acceptable sheet resistance (<15ohm/square) and a scalable
deposition route is ITO.
The relatively high sheet resistance of ITO requires that modules are formed
of multiple strips of photoactive material connected in series. Roll-to-roll print-
ing techniques can reliably ensure electrode spacing of 1–2 mm between strips,
and strip widths of ∼15 mm are empirically found to optimise between the fill
factor and light-harvesting efficiency. For this reason, lab-scale devices with area
>1 cm × 1 cm are considered to give a reasonable indication of the achievable per-
formance of a larger-area module. This architecture has disadvantages, however,
since the intercell gap represents dead module area and shading of a single strip can
lead to a significant drop in module performance unless additional bypass diodes
are used.
Because solution-processed ITO of sufficient quality is not currently available,
ITO is typically pre-sputtered onto PET film and the composite used as the web
for R2R printing. Such an approach has significant drawbacks since indium is
expensive and sputtering is a relatively costly deposition process; ITO/PET is
currently estimated to comprise half of the cost of large-area OPV manufacture.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch08 page 400
400 D. Credgington
8.6 Conclusions
In this chapter we have traced the progress of organic photovoltaics from their early
development, through the various routes found to increase their performance, to
the state-of-the-art. We have set out the current scientific understanding and engi-
neering progress in the use of organic semiconductors in thin-film solar cells,
and highlighted both where there is broad consensus on the electronic processes
involved in light absorption, charge generation, transport and collection, and where
uncertainty still remains. This growing understanding and continued technologi-
cal achievement indicate that excellent performance is achievable (towards that
of current crystalline silicon technology). This will require continued engineering
development to achieve long-lasting, highly efficient, mass-produced devices, but
recent rapid advances give cause for great optimism. The most promising outlook
comes from the sheer size of the space still to explore. Crystalline silicon has been
a stable materials platform for four decades, yet the main driver of increased OPV
performance has been the continual discovery of new organic semiconductors.
There is no indication yet that the ideal OPV materials system has been found, and
until it has, the process of true engineering optimisation remains ahead. Twenty
years from their first discovery, there is every reason to believe that organic pho-
tovoltaics have a bright future.
Acknowledgement
Compiling this chapter would not have been possible without many fruitful dis-
cussions between myself and Professor Sir Richard Friend, who has provided
invaluable insights from his long experience working in the organic electronics
field, and whom I acknowledge unreservedly.
References
Aernouts T., Vanlaeke P., Geens W., Poortmans J., Heremans P., Borghs S., Mertens R.,
Andriessen R. and Leenders L. (2004), ‘Printable anodes for flexible organic solar cell
modules’, Thin Solid Films 451, 22–25.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch08 page 401
Albrecht S., Schindler W., Kurpiers J., Kniepert J., Blakesley J. C., Dumsch I., Allard S.,
Fostiropoulos K., Scherf U. and Neher D. (2012), ‘On the field dependence of free
charge carrier generation and recombination in blends of pcpdtbt/PC70BM: influence
of solvent additives’, J. Phys. Chem. Lett. 3, 640–645.
Andersson L. M., Müller C., Badada B. H., Zhang F., Würfel U. and Inganäs O. (2011),
‘Mobility and fill factor correlation in geminate recombination limited solar cells’,
J. Appl. Phys. 110, 024509.
Bakulin A. A., Rao A., Pavelyev V. G., van Loosdrecht P. H., Pshenichnikov M. S.,
Niedzialek D., Cornil J., Beljonne D. and Friend R. H. (2012), ‘The role of driv-
ing energy and delocalized states for charge separation in organic semiconductors’,
Science 335, 1340–1344.
Ballantyne A. M., Chen L. C., Nelson J., Bradley D. D. C., Astuti Y., Maurano A., Shuttle
C. G., Durrant J. R., Heeney M., Duffy W. and McCulloch I. (2007), ‘Studies of highly
regioregular poly(3-hexylselenophene) for photovoltaic applications’, Adv. Mater. 19,
4544–4547.
Bässler H. (1993), ‘Charge transport in disordered organic photoconductors a Monte Carlo
simulation study’, Phys. Status Solidi B 175, 15–56.
Bässler H. and Kohler A. (2012), ‘Charge transport in organic semiconductors’, Top. Curr.
Chem. 312, 1–65.
Beek W. J. E., Wienk M. M. and Janssen R. A. J. (2004), ‘Efficient hybrid solar cells from
zinc oxide nanoparticles and a conjugated polymer’, Adv. Mater. 16, 1009–1013.
Blakesley J. C. and Neher D. (2011), ‘Relationship between energetic disorder and open-
circuit voltage in bulk heterojunction organic solar cells’, Phys. Rev. B 84, 075210.
Brabec C. J., Shaheen S. E., Winder C., Sariciftci N. S. and Denk P. (2002), ‘Effect of
LiF/metal electrodes on the performance of plastic solar cells’, App. Phys. Lett. 80,
1288–1290.
Bradley D. D. C. (1987), ‘Precursor-route poly( p-phenylenevinylene): Polymer character-
isation and control of electronic properties’, J. Phys. D: Appl. Phys. 20, 1389–1410.
Braun C. L. (1984), ‘Electric field assisted dissociation of charge transfer states as a mech-
anism of photocarrier production’, J. Chem. Phys. 80, 4157–4161.
Brédas J.-L., Norton J. E., Cornil J. and Coropceanu V. (2009), ‘Molecular understanding
of organic solar cells: the challenges’, Acc. Chem. Res. 42, 1691–1699.
Bronstein H., Chen Z., Ashraf R. S., Zhang W., Du J., Durrant J. R., Shakya Tuladhar
P., Song K., Watkins S. E., Geerts Y., Wienk M. M., Janssen R. A. J., Anthopoulos
T., Sirringhaus H., Heeney M. and McCulloch I. (2011), ‘Thieno[3,2-b]thiophene–
diketopyrrolopyrrole-containing polymers for high-performance organic field-effect
transistors and organic photovoltaic devices’, J. Amer. Chem. Soc. 133, 3272–3275.
Cacialli F. and Bruschi P. (1996), ‘Site-selective chemical-vapor-deposition of submicron-
wide conducting polypyrrole films: morphological investigations with the scanning
electron and the atomic force microscope’, J. Appl. Phys. 80, 70–75.
Campoy-Quiles M., Ferenczi T., Agostinelli T., Etchegoin P. G., Kim Y., Anthopoulos T.
D., Stavrinou P. N., Bradley D. D. C. and Nelson J. (2008), ‘Morphology evolution
via self-organization and lateral and vertical diffusion in polymer:fullerene solar cell
blends’, Nat. Mater. 7, 158–164.
Campoy-Quiles M., Kanai Y., El-Basaty A., Sakai H. and Murata H. (2009), ‘Ternary mixing:
a simple method to tailor the morphology of organic solar cells’, Org. Electronics 10,
1120–1132.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch08 page 402
402 D. Credgington
Carsten B., Szarko J. M., Son H. J., Wang W., Lu L., He F., Rolczynski B. S., Lou S. J.,
Chen L. X. and Yu L. (2011), ‘Examining the effect of the dipole moment on charge
separation in donor–acceptor polymers for organic photovoltaic applications’, J. Amer.
Chem. Soc. 133, 20468–20475.
Chambon S., Rivaton A., Gardette J.-L., Firon M. and Lutsen L. (2007), ‘Aging of a donor
conjugated polymer: photochemical studies of the degradation of poly[2-methoxy-5-
(3 , 7 -dimethyloctyloxy)-1,4-phenylenevinylene]’, J. Polymer Sci. Part A: Polymer
Chem. 45, 317–331.
Chappell J., Lidzey D. G., Jukes P. C., Higgins A. M., Thompson R. L., O’Connor S., Grizzi
I., Fletcher R., O’Brien J., Geoghegan M. and Jones R. A. L. (2003), ‘Correlating
structure with fluorescence emission in phase-separated conjugated-polymer blends’,
Nat. Mater. 2, 616–621.
Charas A., Alves H., Alcacer L. and Morgado J. (2006), ‘Use of cross-linkable polyfluorene
in the fabrication of multilayer polyfluorene-based light-emitting diodes with improved
efficiency’, Appl. Phys. Lett. 89, 143519.
Chirvase D., Parisi J., Hummelen J. C. and Dyakonov V. (2004), ‘Influence of nanomor-
phology on the photovoltaic action of polymer-fullerene composites’, Nanotechnology
15, 1317.
Chou S. Y., Krauss P. R. and Renstrom P. J. (1996), ‘Imprint lithography with 25-nanometer
resolution’, Science 272, 85–87.
Clark J., Chang J. F., Spano F. C., Friend R. H. and Silva C. (2009), ‘Determining exci-
ton bandwidth and film microstructure in polythiophene films using linear absorption
spectroscopy’, Appl. Phys. Lett. 94, 163306.
Clarke T. M. and Durrant J. R. (2010), ‘Charge photogeneration in organic solar cells’,
Chem. Rev. 110, 6736–6767.
Credgington D. and Durrant J. R. (2012), ‘Insights from transient optoelectronic analyses
on the open-circuit voltage of organic solar cells’, J. Phys. Chem. Lett. 3, 1465–1478.
Credgington D., Fenwick O., Charas A., Morgado J., Suhling K. and Cacialli F. (2010),
‘High-resolution scanning near-field optical lithography of conjugated polymers’, Adv.
Funct. Mater. 20, 2842–2847.
Credgington D., Jamieson F. C., Walker B., Nguyen T.-Q. and Durrant J. R. (2012), ‘Quantifi-
cation of geminate and non-geminate recombination losses within a solution-processed
small-molecule bulk heterojunction solar cell’, Adv. Mater. 24, 2135–2141.
Crispin A., Crispin X., Fahlman M., Berggren M. and Salaneck W. R. (2006), ‘Transition
between energy level alignment regimes at a low band gap polymer-electrode inter-
faces’, Appl. Phys. Lett. 89, 213503.
de Jong M. P., van Ijzendoorn L. J. and de Voigt M. J. A. (2000), ‘Stability of the inter-
face between indium-tin-oxide and poly(3,4-ethylenedioxythiophene)/poly (styrene-
sulfonate) in polymer light-emitting diodes’, Appl. Phys. Lett. 77, 2255–2257.
Deibel C. and Dyakonov V. (2010), ‘Polymer:fullerene bulk heterojunction solar cells’, Rep.
Progr. Phys. 73, 096401.
Delacote G. M., Fillard J. P. and Marco F. J. (1964), ‘Electron injection in thin films of
copper phtalocyanine’, Solid State Commun. 2, 373–376.
Dennler G., Prall H.-J., Koeppe R., Egginger M., Autengruber R. and Sariciftci N. S.
(2006), ‘Enhanced spectral coverage in tandem organic solar cells’, Appl. Phys. Lett.
89, 073502.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch08 page 403
404 D. Credgington
Frenkel J. (1931), ‘On the transformation of light into heat in solids. I’, Phys. Rev. 37, 17–44.
Gélinas S., Paré-Labrosse O., Brosseau C.-N., Albert-Seifried S., McNeill C. R., Kirov
K. R., Howard I. A., Leonelli R., Friend R. H. and Silva C. (2011), ‘The binding
energy of charge-transfer excitons localized at polymeric semiconductor heterojunc-
tions’, J. Phys. Chem. C 115, 7114–7119.
Giebink N. C., Wiederrecht G. P., Wasielewski M. R. and Forrest S. R. (2011), ‘Thermody-
namic efficiency limit of excitonic solar cells’, Phys. Rev. B 83, 195326.
Glatthaar M., Riede M., Keegan N., Sylvester-Hvid K., Zimmermann B., Niggemann M.,
Hinsch A. and Gombert A. (2007), ‘Efficiency limiting factors of organic bulk het-
erojunction solar cells identified by electrical impedance spectroscopy’, Solar Energy
Mater. Solar Cells 91, 390–393.
Gonzalez-Rabade A., Morteani A. C. and Friend R. H. (2009), ‘Correlation of heterojunction
luminescence quenching and photocurrent in polymer-blend photovoltaic diodes’, Adv.
Mater. 21, 3924–3927.
Grancini G., Maiuri M., Fazzi D., Petrozza A., Egelhaaf H. J., Brida D., Cerullo G. and
Lanzani G. (2013), ‘Hot exciton dissociation in polymer solar cells’, Nat. Mater. 12,
29–33.
Granström M. (1998), ‘Micropatterned luminescent polymer films’, Acta Polymerica 49,
514–517.
Green M. A., Emery K., Hishikawa Y., Warta W. and Dunlop E. D. (2013), ‘Solar cell
efficiency tables (version 41)’, Progr. Photovoltaics 21, 1–11.
Greenham N. C., Peng X. and Alivisatos A. P. (1996), ‘Charge separation and transport in
conjugated-polymer/semiconductor-nanocrystal composites studied by photolumines-
cence quenching and photoconductivity’, Phys. Rev. B 54, 17628–17637.
Grisorio R., Allegretta G., Mastrorilli P. and Suranna G. P. (2011), ‘On the degradation pro-
cess involving polyfluorenes and the factors governing their spectral stability’, Macro-
molecules 44, 7977–7986.
Hagfeldt A., Boschloo G., Sun L., Kloo L. and Pettersson H. (2010), ‘Dye-sensitized solar
cells’, Chem. Rev. 110, 6595–6663.
Halls J. J. M., Pichler K., Friend R. H., Moratti S. C. and Holmes A. B. (1996), ‘Exciton
diffusion and dissociation in a poly( p-phenylenevinylene)/C60 heterojunction photo-
voltaic cell’, Appl. Phys. Lett. 68, 3120–3122.
Halls J. J. M., Walsh C. A., Greenham N. C., Marseglia E. A., Friend R. H., Moratti S. C. and
Holmes A. B. (1995), ‘Efficient photodiodes from interpenetrating polymer networks’,
Nature 376, 498–500.
He Z., Zhong C., Su S., Xu M., Wu H. and Cao Y. (2012), ‘Enhanced power-conversion
efficiency in polymer solar cells using an inverted device structure’, Nat. Photonics 6,
591–595.
Heliatek (2013), ‘Heliatek consolidates its technology leadership by establishing a new
world record for organic solar technology with a cell efficiency of 12%’, 16 January
2013, www.heliatek.com.
Huggins M. L. (1941), ‘Solutions of long chain compounds’, J. Chem. Phys. 9, 440.
Huynh W. U., Dittmer J. J. and Alivisatos A. P. (2002), ‘Hybrid nanorod-polymer solar
cells’, Science 295, 2425–2427.
Irwin M. D., Buchholz B., Hains A. W., Chang R. P. H. and Marks T. J. (2008),
‘P-type semiconducting nickel oxide as an efficiency-enhancing anode interfacial
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch08 page 405
layer in polymer bulk-heterojunction solar cells’, Proc. Natl. Acad. Sci. U S A 105,
2783–2787.
Jamieson F. C., Domingo E. B., McCarthy-Ward T., Heeney M., Stingelin N. and Durrant
J. R. (2012), ‘Fullerene crystallisation as a key driver of charge separation in poly-
mer/fullerene bulk heterojunction solar cells’, Chem. Sci. 3, 485–492.
Jørgensen M., Norrman K., Gevorgyan S. A., Tromholt T., Andreasen B. and Krebs F. C.
(2012), ‘Stability of polymer solar cells’, Adv. Mater. 24, 580–612.
Kallmann H. and Pope M. (1959), ‘Photovoltaic effect in organic crystals’, J. Chem. Phys.
30, 585–586.
Kameoka J., Czaplewski D., Liu H. and Craighead H. G. (2004), ‘Polymeric nanowire
architecture’, J. Mater. Chem. 14, 1503–1505.
Kearns D. and Calvin M. (1958), ‘Photovoltaic effect and photoconductivity in laminated
organic systems’, J. Chem. Phys. 29, 950–951.
Kim J. Y., Kim S. H., Lee H. H., Lee K., Ma W., Gong X. and Heeger A. J. (2006),
‘New architecture for high-efficiency polymer photovoltaic cells using solution-based
titanium oxide as an optical spacer’, Adv. Mater. 18, 572–576.
Kim Y., Cook S., Tuladhar S. M., Choulis S. A., Nelson J., Durrant J. R., Bradley D. D. C.,
Giles M., McCulloch I., Ha C.-S. and Ree M. (2006), ‘A strong regioregularity effect in
self-organizing conjugated polymer films and high-efficiency polythiophene:fullerene
solar cells’, Nat. Mater. 5, 197–203.
Kirchartz T., Pieters B. E., Kirkpatrick J., Rau U. and Nelson J. (2011), ‘Recombination via
tail states in polythiophene:fullerene solar cells’, Phys. Rev. B 83, 115209.
Kniepert J., Schubert M., Blakesley J. C. and Neher D. (2011), ‘Photogeneration and recom-
bination in P3HT/PCBM solar cells probed by time-delayed collection field experi-
ments’, J. Phys. Chem. Lett. 2, 700–705.
Kohn P., Rong Z., Scherer K. H., Sepe A., Sommer M., Müller-Buschbaum P., Friend R. H.,
Steiner U. and Hüttner S. (2013), ‘Crystallization-induced 10-nm structure formation
in P3HT/PCBM blends’, Macromolecules 46, 4002–4013.
Koster L. J. A., Mihailetchi V. D., Ramaker R. and Blom P. W. M. (2005), ‘Light intensity
dependence of open-circuit voltage of polymer:fullerene solar cells’, Appl. Phys. Lett.
86, 123509.
Krebs F. C. (2009), ‘Fabrication and processing of polymer solar cells: a review of printing
and coating techniques’, Solar Energy Mater. Solar Cells 93, 394–412.
Krebs F. C., Fyenbo J. and Jorgensen M. (2010), ‘Product integration of compact roll-to-roll
processed polymer solar cell modules: methods and manufacture using flexographic
printing, slot-die coating and rotary screen printing’, J. Mater. Chem. 20, 8994–9001.
Krebs F. C., Gevorgyan S. A. and Alstrup J. (2009), ‘A roll-to-roll process to flexible polymer
solar cells: model studies, manufacture and operational stability studies’, J. Mater.
Chem. 19, 5442–5451.
Krebs F. C. and Spanggaard H. (2005), ‘Significant improvement of polymer solar cell
stability’, Chemistry of Materials 17, 5235–5237.
Kumar S. and Nann T. (2004), ‘First solar cells based on CdTe nanoparticle/MEH-PPV
composites’, J. Mater. Res. 19, 1990–1994.
Lee J., Vandewal K., Yost S. R., Bahlke M. E., Goris L., Baldo M. A., Manca J. V. and
Voorhis T. V. (2010), ‘Charge transfer state versus hot exciton dissociation in polymer-
fullerene blended solar cells’, J. Amer. Chem. Soc. 132, 11878–11880.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch08 page 406
406 D. Credgington
Lee J. K., Ma W. L., Brabec C. J., Yuen J., Moon J. S., Kim J. Y., Lee K., Bazan G. C.
and Heeger A. J. (2008), ‘Processing additives for improved efficiency from bulk
heterojunction solar cells’, J. Amer. Chem. Soc. 130, 3619–3623.
Lee M. J., Gupta D., Zhao N., Heeney M., McCulloch I. and Sirringhaus H. (2011),
‘Anisotropy of charge transport in a uniaxially aligned and chain-extended, high-
mobility, conjugated polymer semiconductor’, Adv. Funct. Mater. 21, 932–940.
Li G., Shrotriya V., Yao Y. and Yang Y. (2005), ‘Investigation of annealing effects and film
thickness dependence of polymer solar cells based on poly(3-hexylthiophene)’, J. Appl.
Phys. 98, 043704.
Li G., Yao Y., Yang H., Shrotriya V., Yang G. and Yang Y. (2007), “‘Solvent annealing”
effect in polymer solar cells based on poly(3-hexylthiophene) and methanofullerenes’,
Adv. Funct. Mater. 17, 1636–1644.
Li J., Kim S., Edington S., Nedy J., Cho S., Lee K., Heeger A. J., Gupta M. C. and Yates Jr J. T.
(2011), ‘A study of stabilization of P3HT/PCBM organic solar cells by photochemical
active TiOx layer’, Solar Energy Mater. Solar Cells 95, 1123–1130.
Li J., Zhao Y., Tan H. S., Guo Y., Di C.-A., Yu G., Liu Y., Lin M., Lim S. H., Zhou Y., Su
H. and Ong B. S. (2012), ‘A stable solution-processed polymer semiconductor with
record high-mobility for printed transistors’, Sci. Rep. 2, 754.
Li W., Furlan A., Hendriks K. H., Wienk M. M. and Janssen R. A. J. (2013), ‘Efficient
tandem and triple-junction polymer solar cells’, J. Amer. Chem. Soc. 135, 5529–5532.
Lin C.-F., Liu S.-W., Lee C.-C., Hunag J.-C., Su W.-C., Chiu T.-L., Chen C.-T. and Lee J.-H.
(2012), ‘Open-circuit voltage and efficiency improvement of subphthalocyanine-based
organic photovoltaic device through deposition rate control’, Solar Energy Mater. Solar
Cells 103, 69–75.
Linares M., Beljonne D., Cornil J. r. m., Lancaster K., Bré das J.-L., Verlaak S., Mityashin
A., Heremans P., Fuchs A., Lennartz C., Idé J., Méreau R. L., Aurel P., Ducasse L. and
Castet F. (2010), ‘On the interface dipole at the pentacene–fullerene heterojunction: a
theoretical study’, J. Phys. Chem. C 114, 3215–3224.
Liu B., Png R.-Q., Zhao L.-H., Chua L.-L., Friend R. H. and Ho P. K. H. (2012), ‘High
internal quantum efficiency in fullerene solar cells based on crosslinked polymer donor
networks’, Nature Commun. 3, 1321.
Liu C.-Y., Holman Z. C. and Kortshagen U. R. (2008), ‘Hybrid solar cells from P3HT and
silicon nanocrystals’, Nano Letters 9, 449–452.
Liu J., Shi Y. J. and Yang Y. (2001), ‘Solvation-induced morphology effects on the perfor-
mance of polymer-based photovoltaic devices’, Adv. Funct. Mater. 11, 420–424.
Lüssem B., Riede M. and Leo K. (2013), ‘Doping of organic semiconductors’, Phys. Stat.
Solidi A 210, 9–43.
Ma W., Yang C., Gong X., Lee K. and Heeger A. J. (2005), ‘Thermally stable, efficient poly-
mer solar cells with nanoscale control of the interpenetrating network morphology’,
Adv. Funct. Mater. 15, 1617–1622.
MacKenzie R. C. I., Shuttle C. G., Chabinyc M. L. and Nelson J. (2012), ‘Extracting micro-
scopic device parameters from transient photocurrent measurements of P3HT:PCBM
solar cells’, Adv. Energy Mater. 2, 662–669.
Manceau M., Bundgaard E., Carle J. E., Hagemann O., Helgesen M., Sondergaard R.,
Jorgensen M. and Krebs F. C. (2011), ‘Photochemical stability of small π-conjugated
polymers for polymer solar cells: a rule of thumb’, J. Mater. Chem. 21, 4132–4141.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch08 page 407
Manceau M., Chambon S., Rivaton A., Gardette J.-L., Guillerez S. and Lemaître N. (2010a),
‘Effects of long-term uv-visible light irradiation in the absence of oxygen on P3HT
and P3HT:PCBM blend’, Solar Energy Mater. Solar Cells 94, 1572–1577.
Manceau M., Helgesen M. and Krebs F. C. (2010b), ‘Thermo-cleavable polymers: materials
with enhanced photochemical stability’, Polym. Degrad. Stabil. 95, 2666–2669.
Manceau M., Rivaton A., Gardette J.-L., Guillerez S. and Lemaître N. (2009), ‘The mecha-
nism of photo- and therMoOxidation of poly(3-hexylthiophene) (P3HT) reconsidered’,
Polym. Degrad. Stabil. 94, 898–907.
Mandoc M. M., Veurman W., Koster L. J. A., de Boer B. and Blom P. W. M. (2007), ‘Origin
of the reduced fill factor and photocurrent in MDMO-PPV:PCNEPV all-polymer solar
cells’, Adv. Funct. Mater. 17, 2167–2173.
Marsh R. A., Hodgkiss J. M. and Friend R. H. (2010), ‘Direct measurement of electric
field-assisted charge separation in polymer:fullerene photovoltaic diodes’, Adv. Mater.
22, 3672–3676.
Martens H. C. F., Brom H. B. and Blom P. W. M. (1999), ‘Frequency-dependent electrical
response of holes in poly( p-phenylene vinylene)’, Phys. Rev. B 60, R8489–R8492.
Meilanov I. S., Benderskii V. A. and Bliumenfel’d L. A. (1970a), ‘Photoelectric properties
of chlorophyll a and b layers. I. Photocurrents during constant illumination’, Biofizika
15, 822–827.
Meilanov I. S., Benderskii V. A. and Bliumenfel’d L. A. (1970b), ‘Photoelectric properties
of chlorophyll a and b layers. II. Photocurrents during impulse illumination’, Biofizika
15, 958–964.
Mihailetchi V. D., Blom P. W. M., Hummelen J. C. and Rispens M. T. (2003), ‘Cathode
dependence of the open-circuit voltage of polymer:fullerene bulk heterojunction solar
cells’, J. Appl. Phys. 94, 6849–6854.
Mihailetchi V. D., Koster L. J. A. and Blom P. W. M. (2004), ‘Effect of metal electrodes
on the performance of polymer:fullerene bulk heterojunction solar cells’, Appl. Phys.
Lett. 85, 970–972.
Mihailetchi V. D., Wildeman J. and Blom P. W. M. (2005), ‘Space-charge limited photocur-
rent’, Phys. Rev. Lett. 94, 126602.
Miller A. and Abrahams E. (1960), ‘Impurity conduction at low concentrations’, Phys. Rev.
120, 745–755.
Morana M., Azimi H., Dennler G., Egelhaaf H.-J., Scharber M., Forberich K., Hauch J.,
Gaudiana R., Waller D., Zhu Z., Hingerl K., van Bavel S. S., Loos J. and Brabec C. J.
(2010), ‘Nanomorphology and charge generation in bulk heterojunctions based on low-
bandgap dithiophene polymers with different bridging atoms’, Adv. Funct. Mater. 20,
1180–1188.
Moulé A. J. and Meerholz K. (2008), ‘Controlling morphology in polymer–fullerene mix-
tures’, Adv. Mater. 20, 240–245.
Mühlbacher D., Scharber M., Morana M., Zhu Z., Waller D., Gaudiana R. and Brabec C.
(2006), ‘High photovoltaic performance of a low-bandgap polymer’, Adv. Mater. 18,
2884–2889.
Najari A., Berrouard P., Ottone C., Boivin M., Zou Y., Gendron D., Caron W.-O., Legros
P., Allen C. N., Sadki S. and Leclerc M. (2012), ‘High open-circuit voltage solar cells
based on new thieno[3,4-c]pyrrole-4,6-dione and 2,7-carbazole copolymers’, Macro-
molecules 45, 1833–1838.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch08 page 408
408 D. Credgington
Reese M. O., Gevorgyan S. A., Jørgensen M., Bundgaard E., Kurtz S. R., Ginley D. S.,
Olson D. C., Lloyd M. T., Morvillo P., Katz E. A., Elschner A., Haillant O., Currier T.
R., Shrotriya V., Hermenau M., Riede M., R. Kirov K., Trimmel G., Rath T., Inganäs
O., Zhang F., Andersson M., Tvingstedt K., Lira-Cantu M., Laird D., McGuiness C.,
Gowrisanker S., Pannone M., Xiao M., Hauch J., Steim R., DeLongchamp D. M.,
Rösch R., Hoppe H., Espinosa N., Urbina A., Yaman-Uzunoglu G., Bonekamp J.-B.,
van Breemen A. J. J. M., Girotto C., Voroshazi E. and Krebs F. C. (2011), ‘Consensus
stability testing protocols for organic photovoltaic materials and devices’, Solar Energy
Mater. Solar Cells 95, 1253–1267.
Reese M. O., Nardes A. M., Rupert B. L., Larsen R. E., Olson D. C., Lloyd M. T., Shaheen
S. E., Ginley D. S., Rumbles G. and Kopidakis N. (2010), ‘Photoinduced degradation
of polymer and polymer–fullerene active layers: experiment and theory’, Adv. Funct.
Mater. 20, 3476–3483.
Riede M., Uhrich C., Widmer J., Timmreck R., Wynands D., Schwartz G., Gnehr W.-M.,
Hildebrandt D., Weiss A., Hwang J., Sundarraj S., Erk P., Pfeiffer M. and Leo K.
(2011), ‘Efficient organic tandem solar cells based on small molecules’, Adv. Funct.
Mater. 21, 3019–3028.
Rogers J. A., Bao Z., Makhija A. and Braun P. (1999), ‘Printing process suitable for reel-to-
reel production of high-performance organic transistors and circuits’, Adv. Mater. 11,
741–745.
Rogers J. A., Bao Z. and Raju V. R. (1998), ‘Nonphotolithographic fabrication of organic
transistors with micron feature sizes’, Appl. Phys. Lett. 72, 2716–2718.
Sariciftci N. S., Smilowitz L., Heeger A. J. and Wudl F. (1992), ‘Photoinduced electron
transfer from a conducting polymer to buckminsterfullerene’, Science 258, 1474–1476.
Scharber M. C., Mühlbacher D., Koppe M., Denk P., Waldauf C., Heeger A. J. and Brabec
C. J. (2006), ‘Design rules for donors in bulk-heterojunction solar cells — towards
10% energy-conversion efficiency’, Adv. Mater. 18, 789–794.
Segalman R. A., McCulloch B., Kirmayer S. and Urban J. J. (2009), ‘Block copolymers for
organic optoelectronics’, Macromolecules 42, 9205–9216.
Shaheen S. E., Brabec C. J., Sariciftci N. S., Padinger F., Fromherz T. and Hummelen J. C.
(2001), ‘2.5% efficient organic plastic solar cells’, Appl. Phys. Lett. 78, 841–843.
Shoaee S., Clarke T. M., Huang C., Barlow S., Marder S. R., Heeney M., McCulloch I.
and Durrant J. R. (2010), ‘Acceptor energy level control of charge photogeneration in
organic donor/acceptor blends’, J. Amer. Chem. Soc. 132, 12919–12926.
Shoaee S., Subramaniyan S., Xin H., Keiderling C., Tuladhar P. S., Jamieson F., Jenekhe S.
A. and Durrant J. R. (2013), ‘Charge photogeneration for a series of thiazolo-thiazole
donor polymers blended with the fullerene electron acceptors PCBM and ICBA’, Adv.
Funct. Mater. 23, 3286–3298.
Shuttle C. G., Hamilton R., O’Regan B. C., Nelson J. and Durrant J. R. (2010), ‘Charge-
density-based analysis of the current–voltage response of polythiophene/fullerene pho-
tovoltaic devices’, Proc. Natl. Acad. Sci. USA 107, 16448–16452.
Silinsh E. A., Belkind A. I., Balode D. R., Biseniece A. J., Grechov V. V., Taure L. F.,
Kurik M. V., Vertzymacha J. I. and Bok I. (1974), ‘Photoelectrical properties, energy
level spectra, and photogeneration mechanisms of pentacene’, Phys. Stat. Solidi A 25,
339–347.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch08 page 410
410 D. Credgington
Singh S., Jones W. J., Siebrand W., Stoicheff B. P. and Schneider W. G. (1965), ‘Laser
generation of excitons and fluorescence in anthracene crystals’, J. Chem. Phys. 42,
330–342.
Smith M. B. and Michl J. (2010), ‘Singlet fission’, Chem. Rev. 110, 6891–6936.
Son H. J., Wang W., Xu T., Liang Y., Wu Y., Li G. and Yu L. (2011), ‘Synthesis of flu-
orinated polythienothiophene-co-benzodithiophenes and effect of fluorination on the
photovoltaic properties’, J. Amer. Chem. Soc. 133, 1885–1894.
Stabile R., Camposeo A., Persano L., Tavazzi S., Cingolani R. and Pisignano D. (2007),
‘Organic-based distributed feedback lasers by direct electron-beam lithography on
conjugated polymers’, Appl. Phys. Lett. 91, 101110.
Steim R., Kogler F. R. and Brabec C. J. (2010), ‘Interface materials for organic solar cells’,
J. Mater. Chem. 20, 2499–2512.
Street R. A., Cowan S. and Heeger A. J. (2010), ‘Experimental test for geminate recombi-
nation applied to organic solar cells’, Phys. Rev. B 82, 121301.
Stubinger T. and Brutting W. (2001), ‘Exciton diffusion and optical interference in organic
donor–acceptor photovoltaic cells’, J. Appl. Phys. 90, 3632–3641.
Sullivan P., Schumann S., Da Campo R., Howells T., Duraud A., Shipman M., Hatton R.
A. and Jones T. S. (2013), ‘Ultra-high voltage multijunction organic solar cells for
low-power electronic applications’, Adv. Energy Mater. 3, 239–244.
Sun B., Marx E. and Greenham N. C. (2003), ‘Photovoltaic devices using blends of branched
CdSe nanoparticles and conjugated polymers’, Nano Letters 3, 961–963.
Sun Y., Welch G. C., Leong W. L., Takacs C. J., Bazan G. C. and Heeger A. J. (2012),
‘Solution-processed small-molecule solar cells with 6.7% efficiency’, Nat. Mater. 11,
44–48.
Tang C. W. (1986), ‘Two-layer organic photovoltaic cell’, Appl. Phys. Lett. 48, 183–185.
Tang C. W. and Albrecht A. C. (1975), ‘Photovoltaic effects of metal–chlorophyll-a–metal
sandwich cells’, J. Chem. Phys. 62, 2139–2149.
Tengstedt C., Osikowicz W., Salaneck W. R., Parker I. D., Hsu C.-H. and Fahlman M. (2006),
‘Fermi-level pinning at conjugated polymer interfaces’, Appl. Phys. Lett. 88, 053502.
Timmreck R., Olthof S., Leo K. and Riede M. K. (2010), ‘Highly doped layers as efficient
electron-hole recombination contacts for tandem organic solar cells’, J. Appl. Phys.
108, 033108.
Townsend P. D., Pereira C. M., Bradley D. D. C., Horton M. E. and Friend R. H. (1985),
‘Increase in chain conjugation length in highly oriented Durham-route polyacetylene’,
J. Phys. C: Solid State Physics 18, L283–L289.
Treat N. D., Brady M. A., Smith G., Toney M. F., Kramer E. J., Hawker C. J. and Chabinyc M.
L. (2011), ‘Interdiffusion of PCBM and P3HT reveals miscibility in a photovoltaically
active blend’, Adv. Energy Mater. 1, 82–89.
Tributsch H. and Calvin M. (1971), ‘Electrochemistry of excited molecules: photo-
electrochemical reactions of chlorophylls*’, Photochem. Photobiol. 14, 95–112.
Tvingstedt K., Vandewal K., Gadisa A., Zhang F., Manca J. and Inganäs O. (2009), ‘Elec-
troluminescence from charge transfer states in polymer solar cells’, J. Amer. Chem.
Soc. 131, 11819–11824.
Tvingstedt K., Vandewal K., Zhang F. and Inganás O. (2010), ‘On the dissociation efficiency
of charge transfer excitons and Frenkel excitons in organic solar cells: A luminescence
quenching study’, J. Phys. Chem C 114, 21824–21832.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch08 page 411
Vandewal K., Gadisa A., Oosterbaan W. D., Bertho S., Banishoeib F., Van Severen I.,
Lutsen L., Cleij T. J., Vanderzande D. and Manca J. V. (2008), ‘The relation between
open-circuit voltage and the onset of photocurrent generation by charge-transfer absorp-
tion in polymer:fullerene bulk heterojunction solar cells’, Adv. Funct. Mater. 18, 2064–
2070.
Vandewal K., Tvingstedt K., Gadisa A., Inganäs O. and Manca J. V. (2009), ‘On the
origin of the open-circuit voltage of polymer-fullerene solar cells’, Nat. Mater. 8,
904–909.
Veldman D., İpek Ö., Meskers S. C. J., Sweelssen J., Koetse M. M., Veenstra S. C., Kroon
J. M., van Bavel S. S., Loos J. and Janssen R. A. J. (2008), ‘Compositional and electric
field dependence of the dissociation of charge transfer excitons in alternating polyflu-
orene copolymer/fullerene blends’, J. Amer. Chem. Soc. 130, 7721–7735.
Vitoratos E., Sakkopoulos S., Dalas E., Paliatsas N., Karageorgopoulos D., Petraki F., Ken-
nou S. and Choulis S. A. (2009), ‘Thermal degradation mechanisms of PEDOT:PSS’,
Org. Electronics 10, 61–66.
Vos A. D. (1980), ‘Detailed balance limit of the efficiency of tandem solar cells’, J. Phys.
D: Appl. Phys. 13, 839–846.
Wannier G. H. (1937), ‘The structure of electronic excitation levels in insulating crystals’,
Phys. Rev. 52, 191–197.
Watt A., Blake D., Warner J. H., Thomsen E. A., Tavenner E. L., Rubinsztein-Dunlop H.
and Meredith P. (2005), ‘Lead sulfide nanocrystal: conducting polymer solar cells’,
J. Phys. D: Appl. Phys. 38, 2006–2012.
Wetzelaer G. A. H., Koster L. J. A. and Blom P. W. M. (2011a), ‘Validity of the Einstein
relation in disordered organic semiconductors’, Phys. Rev. Lett. 107, 066605.
Wetzelaer G. A. H., Kuik M., Lenes M. and Blom P. W. M. (2011b), ‘Origin of the dark-
current ideality factor in polymer:fullerene bulk heterojunction solar cells’, Appl. Phys.
Lett. 99, 153506.
Wilson M. W. B., Rao A., Clark J., Kumar R. S. S., Brida D., Cerullo G. and Friend R. H.
(2011), ‘Ultrafast dynamics of exciton fission in polycrystalline pentacene’, J. Amer.
Chem. Soc. 133, 11830–11833.
Xin H., Reid O. G., Ren G., Kim F. S., Ginger D. S. and Jenekhe S. A. (2010), ‘Poly-
mer nanowire/fullerene bulk heterojunction solar cells: how nanostructure determines
photovoltaic properties’, ACS Nano 4, 1861–1872.
Yu G., Gao J., Hummelen J. C., Wudl F. and Heeger A. J. (1995), ‘Polymer photovoltaic
cells: enhanced efficiencies via a network of internal donor-acceptor heterojunctions’,
Science 270, 1789–1791.
Yun M., Myung N. V., Vasquez R. P., Lee C., Menke E. and Penner R. M. (2004), Electro-
chemically grown wires for individually addressable sensor arrays’, Nano Letters 4,
419–422.
Zhong H., Li Z., Deledalle F., Fregoso E. C., Shahid M., Fei Z., Nielsen C. B., Yaacobi-Gross
N., Rossbauer S., Anthopoulos T. D., Durrant J. R. and Heeney M. (2013), ‘Fused
dithienogermolodithiophene low band gap polymers for high-performance organic
solar cells without processing additives’, J. Amer. Chem. Soc. 135, 2040–2043.
Zhou Y., Fuentes-Hernandez C., Shim J., Meyer J., Giordano A. J., Li H., Winget P.,
Papadopoulos T., Cheun H., Kim J., Fenoll M., Dindar A., Haske W., Najafabadi
E., Khan T. M., Sojoudi H., Barlow S., Graham S., Brédas J.-L., Marder S. R., Kahn A.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch08 page 412
412 D. Credgington
CHAPTER 9
and
JAMES R. DURRANT
Department of Chemistry, Imperial College London SW7 2AZ, UK
Materials Science Centre,
College of Engineering, University of Swansea,
Swansea SA12 7Ax, UK
[email protected]
On the arid lands there will spring up industrial colonies without smoke and without
smokestacks; forests of glass tubes will extend over the plains and glass buildings will rise
everywhere; inside of these will take place the photochemical processes that hitherto have
been the guarded secret of the plants, but that will have been mastered by human industry
which will know how to make them bear even more abundant fruit than nature, for nature
is not in a hurry and mankind is. And if in a distant future the supply of coal becomes
completely exhausted, civilisation will not be checked by that, for life and civilisation will
continue as long as the Sun shines!
Giacomo Ciamician, Eighth International Congress of Applied Chemistry,
Washington and New York, September 1912.
9.1 Introduction
Photovoltaic (PV) devices are based on charge separation at an interface of two
materials of different conduction type or mechanism. To date, this field has been
dominated by solid-state junction devices, usually made of silicon, and profiting
from the experience and material availability resulting from the semiconductor
industry. The dominance of the photovoltaic field by inorganic solid-state junction
413
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch09 page 414
devices is now being challenged by the emergence of a range of new device con-
cepts, including devices based on nanocrystalline and conducting polymer films.
These devices offer the prospect of very low-cost fabrication and present a range
of attractive features that will facilitate market entry. It is now possible to depart
completely from the classical solid-state junction device by replacing the phase
contacting the semiconductor by an electrolyte, liquid, gel or solid, thereby form-
ing a photoelectrochemical cell. The phenomenal progress realised recently in the
fabrication and characterisation of nanocrystalline materials has opened up vast
new opportunities for these systems. Contrary to expectation, devices based on
interpenetrating networks of mesoscopic semiconductors have shown strikingly
high conversion efficiencies, competing with those of conventional devices. The
prototype of this family of devices is the dye-sensitised solar cell (DSSC), which
realises the optical absorption and charge-separation processes by the association
of a sensitiser as light-absorbing material with a wide-bandgap semiconductor of
nanocrystalline morphology (O’Regan and Grätzel, 1991).
Recently, solid-state embodiments of mesoscopic solar have emerged where
the sensitiser is replaced by a semiconductor quantum dot (Grätzel et al., 2012)
or a perovskite pigment. Solid-state cells based on perovskites as light harvesters
have stunned the PV community by their amazing performance, certified power
conversion efficiencies having reached 14.14% only within one year after their
inception (Burschka et al., 2013).
injection of electrons from photoexcited dye molecules into the conduction band
of the n−type semiconductor substrates (Gerischer and Tributsch, 1968) — dates
only to the 1960s. In the years that followed it became recognised that the dye
could function most efficiently if chemisorbed on the surface of the semiconduc-
tor (Tsubomura et al., 1976; Anderson et al., 1979; Dare-Edwards et al., 1980).
The concept of using dispersed particles to provide a sufficient interface area then
emerged (Dung et al., 1984), and was subsequently employed for photoelectrodes
(Desilvestro et al., 1985).
Titanium dioxide (TiO2 ) quickly became the semiconductor of choice for
the photoelectrode on account of its many advantages for sensitised photochem-
istry and photoelectron chemistry: it is a low-cost, widely available, non-toxic and
biocompatible material, and as such is even used in healthcare products as well as
industrial applications such as paint pigmentation. Initial studies of the TiO2 -based
DSSC employed tris-bipyridyl ruthenium(II) dyes, which are paradigm sensitisers
for photochemical studies, functionalised by the addition of carboxylate groups to
attach the chromophore to the oxide substrate by chemisorption. Progress there-
after was incremental, a synergy of structure, substrate roughness and morphol-
ogy, dye photophysics and electrolyte redox chemistry, until the announcement
in 1991 (O’Regan and Grätzel, 1991) of a sensitised electrochemical photovoltaic
device with an energy conversion efficiency of 7.1% under solar illumination. That
evolution has continued progressively since then, with certified efficiencies now
over 12%. These advances in efficiency have been complemented by significant
advances in processability, cost and stability which have greatly enhanced the
commercial viability of this technology.
light -
I
I3
external circuit
electrons
Figure 9.1 Schematic of a liquid electrolyte dye-sensitised solar cell. Photoexcitation of the
sensitiser dye is followed by electron injection into the conduction band of the mesoporous
oxide semiconductor, and electron transport through the metal oxide film to the TCO-coated,
glass, working electrode. The dye molecule is regenerated by the redox system, which is
itself regenerated at the platinised counter electrode by electrons passed through the external
circuit.
Figure 9.2 Scanning electron micrograph of a typical mesoscopic TiO2 film employed in
DSSCs. Note the bipyramidal shape of the particles, with (101) oriented facets exposed.
The average particle size is 20 nm.
COOH
COO-
N
S
C
N N
Ru
N N
C
S N
COO-
COOH
Figure 9.3 Chemical structure of the N719 ruthenium complex used as a charge-transfer
sensitiser in dye-sensitised solar cells. The N3 dye has the same structure, except that all
four carboxylates are protonated.
effectively screens out any macroscopic electric fields. Charge separation is there-
fore primarily driven by the inherent energetics (oxidation/reduction potentials) of
the different species at the TiO2 /dye/electrolyte interface rather than by the pres-
ence of macroscopic electrostatic potential energy gradients. Similarly, charge-
transport processes are primarily diffusive, driven by concentration gradients in
the device generated by the photoinduced charge separation. Photoinduced charge
separation occurs at the dye-sensitised TiO2 /electrolyte interface. Electron injec-
tion requires the dye excited state to be more reducing than the TiO2 conduc-
tion band edge, enabling transfer of an electron from photoexcited dye into the
metal oxide. This energetic requirement is equivalent to the dye excited-state low-
est unoccupied molecular orbital (LUMO) having a lower electron affinity than
the electrode conduction-band edge. Similarly, regeneration of the dye ground
state by the redox couple requires the dye cation to be more oxidising than the
iodide/triiodide redox couple. Charge separation in DSSCs can be regarded as a
two-step redox cascade, resulting in the injection of electrons into the TiO2 elec-
trode and the concomitant oxidation of the redox electrolyte. Figure 9.4 shows
typical values for the interfacial energetics in DSSCs. Both charge-separation
reactions are thermodynamically downhill and can be achieved with near unity
quantum efficiency in efficient devices.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch09 page 420
conducting sensitiser
glass TiO2 dye electrolyte cathode
U/V
vs. NHE CB injection +
S*/ S
–0.5
EF
maximum
0 voltage
hν
Red /Ox
0.5 diffusion
regeneration
1.0 +
S/S
e
–
e–
Figure 9.4 Energetics of operation of DSSCs. The primary free-energy losses are associ-
ated with electron injection from the excited sensitiser into the TiO2 conduction band and
regeneration of the dye by the redox couple. The voltage output of the device is approx-
imately given by the splitting between the TiO2 Fermi level (dashed line) and the redox
potential of the redox electrolyte.
Following charge separation, charge collection from the device requires trans-
port of the photogenerated charges to their respective electrodes. For an efficient
DSSC under AM 1.5 solar irradiance, these charge fluxes are of the order of
20 mA cm−2 . The high ionic concentrations in the device effectively screen out
any macroscopic electric fields, thereby removing any significant drift compo-
nent of these transport processes. The transport of both electrons and redox ions
is therefore primarily driven by diffusion processes resulting from concentration
gradients. Under optimum conditions (e.g. good TiO2 nanoparticle interconnec-
tions and a low-viscosity electrolyte), these charge-transport processes (electrons
towards the FTO working electrode and the oxidised component of redox couple
towards the counter electrode) can be efficiently driven with only modest con-
centration gradients, and therefore only small free-energy losses (<50 meV). At
the counter electrode, the oxidised redox couple is re-reduced, a platinum catalyst
typically being used to enable this reaction to proceed with minimal overpotential
(again <50 meV). A similarly ohmic contact is achieved at the TiO2 /FTO inter-
face. It is apparent that the energetics at the TiO2 /dye/electrolyte interface are of
primary importance in determining the overall device energetics.
Power output from the DSSC requires not only efficient charge collection by
the electrodes but the generation of a photovoltage, corresponding to a free-energy
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch09 page 421
difference between the working and counter electrodes. In the dark at equilibrium,
the Fermi energy of the TiO2 electrode (corresponding to the free energy of elec-
trons in this film after thermalisation) equilibrates with the mid-point potential
of the redox couple, resulting in zero output voltage. Under these conditions, the
TiO2 Fermi level lies deep within the bandgap of the semiconductor, and the film
is effectively insulating, with a negligible electron density in the TiO2 conduction
band. Photoexcitation results in electron injection into the TiO2 conduction band
and concomitant hole injection into (that is, oxidation of) the redox electrolyte. The
high concentrations of oxidised and reduced redox couple present in the electrolyte
in the dark mean that this photooxidation process does not result in a significant
change in redox potential of the electrolyte, which remains effectively fixed at
its dark, resting value. In contrast, electron injection into the TiO2 conduction
band results a dramatic increase in electron density (from approx. ≤1013 cm−3 to
≥1018 cm−3 ), raising the TiO2 Fermi level towards the conduction-band edge, as
illustrated in Fig. 9.4, and allowing the film to become conducting. This shift of the
TiO2 Fermi level under irradiation increases the free energy of injected electrons
and is responsible for the generation of the photovoltage in the external circuit.
The mid-point potential of the redox couple is given by the Nernst equation,
and is therefore dependent on the relative concentrations of iodide and iodine. The
concentrations of these species required for efficient device function are in turn
constrained by kinetic requirements of dye regeneration at the working electrode,
and iodide regeneration at the counter electrode, as discussed below. Typical con-
centrations of these species for the iodine/iodide couple are in the range 0.1–0.7 M
iodide and 10–200 mM iodine, constraining the mid-point potential of this elec-
trolyte to ∼0.4 V vs. NHE. It should furthermore be noted that in the presence
of excess iodide, the iodine is primarily present in the form I− 3 , resulting in this
electrolyte often being referred to as the iodide/triiodide redox couple.
Determining the energetics of the TiO2 conduction band is more complex. As
with most oxides, the surface of TiO2 may be more or less protonated depending
on the pH of the surrounding medium. The resultant changes in surface charge
cause the surface potential to exhibit a Nernstian dependence on effective pH,
shifting by 60 mV per pH unit (Rothenberger et al., 1992). In bulk metal oxides,
surface charge can result in significant bending of the conduction and valence
bands adjacent to the surface. However, in the mesoporous TiO2 films employed in
DSSCs, the nanoparticles are too small to support significant band bending. As a
consequence, the whole conduction band of such mesoporous films shifts with the
surface potential. At pH 1 in aqueous solutions, the conduction band edge for TiO2
has been reported at −0.06 V vs. NHE, shifting negatively as the pH is increased
(Hengerer et al., 2000). Dye-sensitised solar cells typically employ organic rather
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch09 page 422
G/eV
Dye*
electron injection ps
+ –
Dye / e TiO2
regeneration reaction
1 ~1 µs
decay to ground – –
hν ~ns I3 / e TiO2
recombination to dye*,
µs – ms charge transport to
electrodes, ms – s
recombination to
electrolyte, ms – s
Dye
0
Figure 9.5 State diagram representation of the kinetics of DSSC function. Forward pro-
cesses of light absorption, electron injection, dye regeneration and charge transport are
indicated by red arrows. The competing loss pathways of excited-state decay to ground and
electron recombination with dye cations and oxidised redox couple are shown in black.
The vertical scale corresponds to the free energy G stored in the charge-separated states.
Note the free energy of injected electrons is determined by the Fermi level of the TiO2 ;
the figure is drawn assuming a TiO2 Fermi level 0.8 V above the chemical potential of the
redox electrolyte, corresponding to the Fermi level of a typical iodide/iodine-based DSSC
at open circuit. It should be noted that at maximum power point, the Fermi level is lower
due to charge transport to the electrodes (also shown in diagram).
results in increased trap filling. Typical electron-transport times under solar irra-
diation are of the order of milliseconds (Peter and Wijayantha, 2000; Frank et al.,
2004). If a low-viscosity solvent such as acetonitrile is used, transport of the oxi-
dised redox couple to the counter electrode is not rate-limiting. However, the use
of higher viscosity solvents more suitable for practical device applications (for rea-
sons of device stability) can result in significantly lower ionic diffusion constants,
with the resultant iodide/triiodide concentration gradients causing significant free-
energy (series resistance) losses, and also potentially accelerating interfacial charge
recombination.
Given the relatively slow timescale for charge transport in DSSCs compared
with most other photovoltaic devices, and the extensive interfacial area available
for charge recombination in the device (due to its mesoscopic structure), it is
remarkable that the quantum efficiency of charge collection can approach unity. The
key factor enabling this high efficiency is the slow rate constant for the interfacial
charge recombination of injected electrons with the oxidised redox couple. This
reaction is a multi-electron reaction, most simply being described by the equation
which must therefore proceed via one or more intermediate states. The mechanism
of this reaction has been extensively studied, and while the details remain somewhat
controversial, it is apparent that without a suitable catalyst such as platinum, one or
more of the intermediate steps exhibits a significant activation barrier, resulting in
a slow overall rate constant for this reaction. The low rate constant for this recom-
bination reaction on TiO2 , contrasting to the facile electrochemistry of this redox
couple on the platinised counter electrode, is a key factor behind the remarkable
efficiencies achieved to date for DSSCs. Similarly, interface engineering through
sensitiser design to minimise this recombination rate constant has been key to
recent successes with alternative one electron couples (Feldt et al., 2010). Nev-
ertheless, Eq. (9.1) is the primary recombination pathway in DSSCs. The flux of
this recombination pathway increases with increasing electron density in the TiO2
electrode (and therefore with the TiO2 Fermi level or cell voltage). In the dark it is
responsible for the diode-like leakage current observed in current–voltage scans,
while under illumination it is the primary factor limiting the voltage output of the
device.
The kinetic competition between charge transport and recombination in
DSSCs has been analysed in terms of an effective carrier diffusion length L n ,
given by
L n = Deff τ (9.2)
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch09 page 426
where Deff is the effective electron diffusion coefficient, and τ the electron lifetime
due to the charge-recombination reaction given by Eq. (9.1) (Peter and Wijayantha,
2000). Deff increases with light intensity (due to the increased electron density in
the TiO2 film) whilst τ shows a proportional decrease, resulting in L n being largely
independent of light intensity. Typical values for L n are 5–20 µm, and even longer
near the optimum power point for cells with >10% conversion efficiency, consistent
with the high carrier-collection efficiencies observed in efficient DSSCs.
It is important to emphasise that the energetics and kinetics of DSSC function
are not independent considerations. The kinetics of the interfacial electron-transfer
dynamics depend strongly on the energetics of the TiO2 /dye/electrolyte interface
and on the density of electrons in the TiO2 (i.e. the TiO2 Fermi level). Raising the
energy of the TiO2 conduction band reduces recombination losses (as for a given
TiO2 Fermi level, the electron density in the TiO2 film will be lower), and therefore
may give a high cell output voltage, but at the expense of a lower free energy driving
force for charge separation, which may result in a lower quantum efficiency for
charge generation and therefore a lower output current. In practice, modulation of
these energetics and kinetics to achieve optimum device performance remains one
of the key challenges in DSSC research and development.
2009; Hagfeldt et al., 2010; Mishra et al., 2010; Ning et al., 2010; Walter et al.,
2010; Qin and Peng, 2012; McEvoy and Grätzel, 2013). The attachment group of
the dye ensures that it spontaneously assembles as a molecular layer on exposing
the oxide film to the dye solution. This molecular dispersion ensures a high prob-
ability that, once a photon is absorbed, the excited state of the dye molecule will
relax by electron injection into the semiconductor conduction band. However, the
optical absorption of a single monolayer of dye is weak, a fact which originally was
cited as ruling out the possibility of high-efficiency sensitised devices, as it was
assumed that smooth substrate surfaces would be imperative in order to avoid the
recombination loss mechanism associated with rough or polycrystalline structures
in solid-state photovoltaics. This objection was invalidated by recognising that the
injection process places electrons in the semiconductor lattice, spatially separated
from the positive charge carriers in the electrolyte or a hole carrier. Hence, mas-
tering the interface is key to slow down the rate of charge recombination. The dye
molecules themselves, if judiciously designed, can block the access of the redox
shuttle to the interface providing a barrier for charge recombination. By now, the
use of nanocrystalline thin-film structures with a roughness factor of >1000 has
become standard practice.
Polypyridyl complexes of ruthenium and osmium have over many years main-
tained a lead in photovoltaic performance in terms of both conversion yield and
long-term stability. Initially, interest focused on sensitisers having the general struc-
ture ML2 (X)2 where L stands for 2,2 -bipyridyl-4,4-dicarboxylic acid, M is Ru or
Os, and X represents a halide, cyanide, thiocyanate, acetyl acetonate, thiocarba-
mate or water substituent. In particular, the ruthenium complex cis-RuL2 (NCS)2 ,
known as N3 dye, became the paradigm heterogeneous charge-transfer sensitiser
for mesoporous solar cells (Nazeeruddin et al., 1993). The absorption spectrum of
fully protonated N3 has maxima at 518 and 380 nm, with extinction coefficients of
1.3 × 104 M−1 cm−1 and 1.33 × 104 M−1 cm−1 , respectively. The complex emits
at 750 nm, the excited-state lifetime being 60 ns. The optical transition has MLCT
(metal-to-ligand charge-transfer) character: excitation of the dye involves transfer
of an electron from the metal to the π ∗ orbital of the surface-anchoring carboxy-
lated bipyridyl ligand, from where it is released in a timescale of femtoseconds
to picoseconds into the conduction band of TiO2 , generating electric charges with
near unit quantum yield.
Discovered in 1993, the photovoltaic performance of N3 was unmatched for
eight years by virtually hundreds of other complexes that have been synthesised and
tested. However, in 2001 the ‘black dye’ tri(cyanato)-2,22 -terpyridyl-4,44 -9-
tricarboxylate)Ru(II) achieved 10.4% AM1.5 solar-to-power conversion efficiency
in full sunlight (Nazeeruddin et al., 2001). Conversion efficiencies have meanwhile
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch09 page 428
Figure 9.6 Spectral response of the photocurrent for the N3, black dye and DX1 sensitisers.
The incident photon-to-current conversion efficiency (IPCE) is plotted as a function of
excitation wavelength. The structure of the sensitisers and their short-circuit photocurrent
is shown below the graph.
single-junction converters. From there the IPCE rises gradually until at 700 nm it
reaches a plateau of ∼80%. If one accounts for reflection and absorption losses in
the conducting glass, the conversion of incident photons to electric current is prac-
tically quantitative over the whole visible domain. From the overlap integral of the
curves in Fig. 9.6 with the AM 1.5 solar spectrum, the short-circuit photocurrents
of the N3, black dye and DX1 sensitisers cells are predicted to be 17 mA cm−2,
22 mA cm−2 and 27.6 mA cm−2 under standard AM 1.5 G reporting conditions,
respectively, in agreement with experimental observations.
Thus, over the last decade a whole new family of judiciously engineered ruthe-
nium complexes has emerged that show excellent photovoltaic performance (Wang
et al., 2005a; Reddy et al., 2007; Grätzel, 2009; Kalyanasundaram and Grätzel,
2010). The spectral response of the photocurrent for of the DX1 dye, shown by the
blue curve in Fig. 9.6, illustrates the amazing advance in this area (Kinoshita et al.,
2013). By replacing the three thiocyanate groups in the black dye with two chloride
ions and one phosphine ligand, a new panchromatic sensitiser, coded DX1, was
fashioned that harvests sunlight over the whole visible and near-IR range up to
1000 nm, closely matching the spectral response of silicon photovoltaic cells and
generating a short circuit photocurrent of 27.6 mA cm−2 .
Recently, new donor acceptor porphyrins have been developed, in particular
the YD2-o-C8 zinc complex, whose PCE exceeds that of the black dye when used
in conjunction with Co((II/III)(bipy)3 complexes as a redox couple (Yella et al.,
2011). The Nernst potential of this redox mediator is more positive than that of
the iodide/triiodide couple, yielding substantial gains in open-circuit potential of
the cells. The performance of YD2-o-C8 is enhanced by judicious substitution
of two meso-positions at the tetrapyrrole ring with a donor and acceptor moiety
and the two remaining meso-positions bearing phenyl substituents with two bulky
alkoxy groups in ortho-positions. This increases the light-harvesting capacity of the
porphyrin and blocks the unwanted interfacial electron back transfer. Further gain
in efficiency was obtained via the use of a co-sensitiser, coded Y123, absorbing
strongly in the green spectral region. The structure of the two sensitisers is shown
in Fig. 9.7, which presents a plot of the photocurrent density versus voltage for a
mixture of the two dyes, adsorbed at the surface of the nanocrystalline titania film
and the IPCE spectra for this. The i sc , Voc and fill factor values under standard
air mass (AM 1.5) reporting conditions were 17.8 mA cm−2 , 0.935 V and 0.74,
respectively, yielding a power conversion efficiency of 12.3%.
C6H13O
NC
N N OH
O
S S
N Zn N OH
N N O
C6H13 C6H13
C6H13O
C6H13O
(Voc ), the fill factor (ηfill ) and the standard incident solar irradiance (E so =
1000 W m−2 ) as
The PCE value of 12.3% achieved with the YD2-o-C8/Y123 co-sensitised tita-
nia films present a new record for laboratory cells based on redox electrolytes. How-
ever, solid-state DSSC embodiments based on perovskite light harvesters achieve
currently a PCE of 15% (Burschka et al., 2013). Due to significant industrial upscal-
ing efforts, the conversion efficiency of DSSC modules has also been rapidly rising
over the last few years. Using the black ruthenium dye along with an organic co-
sensitiser, Sony Corporation has reached a certified PCE efficiency under AM
1.5 standard conditions of 9.9% (Green et al., 2011). This implies that scaling up
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch09 page 431
dye-sensitised laboratory cells by a factor of at least 100 does not entail a significant
decrease in photovoltaic performance, the loss being mainly due to interconnects
or collector grids reducing the active device area.
Further substantial gains in the PCE of the DSSC can be expected in the near
term by implementing the substantial increase in the open-circuit photovoltage
that can be achieved with cobalt complexes rather than iodide/triiodide as redox
electrolyte. For example, a very high Voc of 1.1 V was recently reached with the
cobalt bisbipyridyl pyrazol complex (Yum et al., 2012) whose redox potential
matches well that of the sensitiser ground state, reducing the free-energy loss in
the regeneration step from 0.6 eV for I− /I−3 to 0.2 eV, as shown by the energy level
diagram in Fig. 9.8.
Figure 9.8 Energy-level diagram of a DSSC. Shown are the ground and excited-state redox
potentials for a typical sensitiser as well as for the iodide/triiodide redox couple and a
series of cobalt complexes; bpy-pz and py-pz represent bipyridylpyrazol and pyridylpyrazol,
respectively.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch09 page 432
Several publications have dealt with the use of stacked DSSC configurations in
which two dyes absorbing different parts of the solar spectrum serve as sensitisers
(Durr et al., 2004; Kubo et al., 2004). We have demonstrated that a tandem device
comprising a DSSC as a top cell for high-energy photons and a copper indium
gallium selenide (CIGS) thin-film bottom cell, capturing the red and near-IR solar
emission respectively, produces AM 1.5 solar-to-electric conversion efficiencies
greater than 15% (Liska et al., 2006).
Figure 9.9 shows the I –V curve and other characteristics obtained under
AM 1.5 insolation for one of the first embodiments of this two-terminal tandem
device. The performance of the tandem was clearly superior to that of the individ-
ual cells, despite the fact that the short-circuit currents of the two cells were not
perfectly matched, as evidenced by the shoulder in the I –V curve. Likewise, no
effort was made to minimise the optical losses. This leaves no doubt that further
efficiency gains reaching well beyond the 20% mark can be expected from the
fructuous marriage of these two thin-film PV technologies (Liska et al., 2006).
Combining a relatively low-cost thin-film CIGS substrate cell with a DSSC super-
strate cell may be a cheaper method of achieving efficiencies above 15% than use
of a high-efficiency CIGS cell alone.
However, in order to minimise optical and ohmic losses it is mandatory to
integrate the two cells in a monolithic device in which they are series-connected
–14
–12
current density / mA cm–2
–10
–8
–6
–4
–2
2
0.0 0.2 0.4 0.6 0.8 1.0 1.2 1.4 1.6
potential / V
9.4.4 Stability
Unlike amorphous silicon, which suffers from degradation due to the well-known
Staebler–Wronski effect, the intrinsic stability of the DSSC has been confirmed
by extensive accelerated light-soaking tests carried out over the last fifteen years.
One major issue that has been settled during this period is that the sensitisers
employed in the current DSSC embodiments can sustain twenty years of outdoor
service without significant degradation. However, as new and more advanced dye
structures emerge, and in order to avoid repeating these lengthy tests every time
the sensitiser is modified, kinetic criteria have been developed to allow prediction
of long-term sensitiser performance.
Figure 9.10 illustrates the catalytic cycle that the sensitiser undergoes during
cell operation. Critical for stability are side reactions that occur from the excited
state (S*) or the oxidised state of the dye (S+ ), which would compete with electron
injection from the excited dye into the conduction band of the mesoscopic oxide
s h•
regeneration excitation
k reg
s+ s*
k1
2
kinj
product product
decomposition injection decomposition
Figure 9.10 The catalytic cycle of the sensitiser during DSSC operation.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch09 page 434
and with the regeneration of the sensitiser. These destructive channels are assumed
to follow first or pseudo-first-order kinetics and are assigned the rate constants
k1 and k2 . Introducing the two branching ratios P1 = kinj /(k1 + k inj) and P2 =
k reg /(k2 + kreg ), where kinj and kreg are the first-order or pseudo-first-order rate
constants for the injection and regeneration process, respectively, the fraction of
the sensitiser molecules that survives n cycles is given by the product (P1 × P2 )n .
A simple calculation (Grätzel, 2006) shows that the sum of the branching ratios for
the two bleeding channels should not exceed 1 × 10−8 in order for the sensitiser
to turn over 100 million times, corresponding to a lifetime of at least twenty years.
The turnover frequency of the dye in the working DSSC, averaged diurnally and
seasonally, is about 0.16 s−1 .
For most of the common sensitisers, the rate constant for electron injection
from the excited state of the dye to the conduction band of the TiO2 particles
is in the picosecond or femtosecond range. Assuming k inj = 1010 –1013 s−1 , any
destructive side reaction should have k 1 < 102 s−1 . Ruthenium sensitisers of the
N719 or C106 type readily satisfy this condition as the decomposition from the
excited-state level occurs at a much lower rate than the 102 s−1 limit. Precise kinetic
information has also been gathered for the second destructive channel involving
the oxidised state of the sensitiser, the key parameter being the ratio k2 /kreg of the
rate constants for the degradation of the oxidised form of the sensitiser and its
regeneration. The S+ state of the sensitiser can readily be produced by chemical
or electrochemical oxidation and its lifetime can be independently determined
by absorption spectroscopy. A typical value of k 2 is around 10−4 s−1 while the
regeneration rate constant is at least in the 105 s−1 range. Hence the branching ratio
is well below the limit of 10−8 , which can be tolerated to achieve the 100 million
turnovers and a twenty-year lifetime for the sensitiser. Apart from the photoinduced
degradation of the sensitiser during the catalytic cycle, its chemical reaction with
constituents of the electrolyte or its desorption or its mere reorganisation on the
surface to a non-productive conformation could possibly harm cell performance
in the long term. An appropriate way to rule out the existence of any problems of
this type is to carry out accelerated tests under heat stress.
Many long-term tests that have been performed with the N3-type ruthenium
complexes have confirmed the extraordinary stability of these charge-transfer sen-
sitisers. For example, a European consortium supported by the Joule program
(Hinsch et al., 2001) confirmed cell photocurrent stability during 8,500 hours of
light soaking at 2.5 Suns, corresponding to about 56 million turnovers of the dye
without any significant degradation. More recently, DSSC lifetimes of over 20
years have been projected from continuous light-soaking tests performed over
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch09 page 435
20,000 hours (Harakirisun and Desilvestro, 2011). These results corroborate the
projections from the kinetic considerations made above.
A more difficult task has been to achieve stability under prolonged stress at
higher temperatures, i.e. 80–85◦ C. The introduction of hydrophobic sensitisers
has been particularly rewarding in allowing DSSCs to meet, for the first time,
the specifications laid out for outdoor applications of silicon photovoltaic cells
(Wang et al., 2003a). In addition these dyes show enhanced extinction coeffi-
cients owing to the extension of the π-conjugation onto one of the bipy ligands
by styrene moieties. Taking advantage of these properties and using a novel robust
electrolyte formulation, a ≥8% efficient DSSC that shows strikingly stable perfor-
mance under both prolonged thermal stress and light soaking was realised (Wang
et al., 2005b; Sauvage et al., 2011). Dyesol recently announced completion of a
high-temperature test carried out with a new electrolyte formulation based on sul-
fone, a polar non-toxic high boiling point solvent and using the above mentioned
Y123 dye as a sensitiser. Impressively, stable performance was maintained over
3000 hours, establishing beyond any doubt that the DSSC can survive long-term
heat exposure without any significant degradation.
While impressive progress has been made in the development of stable, non-
volatile electrolyte formulations, the efficiencies of these systems are presently
in the 7–10% range (Sauvage et al., 2011), below the 11.9% reached with Ru-
based sensitisers and volatile solvents. Future research efforts will be dedicated
to reducing the performance gap between these systems. The focus will be on
hole conductors and solvent-free electrolytes, in particular ionic liquids. The latter
are a particularly attractive choice for the first commercial modules, owing to
their high stability, negligible vapour pressure and excellent compatibility with the
environment. Fujikura (Arakawa et al., 2009) has announced excellent stability data
with cells based on solid (gel) ‘nanocomposite’ electrolytes based on a mixture
of ionic liquids and silica nanoparticles. Modules showing 8% PCE passed all
accelerated stability tests required by the IEC 61646 protocol for outdoor PV
panels including exposure to 85% humidity at 85◦ C for 1000 hours, long-term light
soaking and cycling the temperature of the modules between −40◦ C and +90◦ C.
These results confirm the extraordinary robustness of dye-sensitised solar cells
when employed in conjunction with appropriately designed electrolytes, showing
their aptness for outdoor deployment.
In keeping with the encouraging results obtained by Fujikura, the Israeli com-
pany 3GSolar announced that its DSSC modules have operated on the company’s
rooftop continuously for two years and they continue to perform to the same stan-
dard as on the day they were placed outdoors (3GSolar, 2011). Importantly, 3GSolar
and ECN (DeWild-Scholten and Veltkamp, 2007) have also performed life-cycle
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch09 page 436
analysis, which shows the energy payback time for the DSSC to be less than one
year in southern European climates as compared with about three years for silicon
solar cells at that time; a more recent analysis shows the latter has improved to
1.3–1.5 years (Mann et al., 2013).
Gold
Glass
OCH3 OCH3
OCH3 OCH3
h+ HH3COCO
3 NN N OCH3 N OCH3
Light H3CO N N OCH3
e-
H3CO OCH3
N OCH3
N OCH3
OCH3 OCH3
Figure 9.11 Cross-sectional view of a solid-state DSSC containing the hole conductor
spiro-OMeTAD, the structure of which is indicated on the right. Left-hand drawing courtesy
of B. O’Regan.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch09 page 437
Sargent, 2012). These simple devices function surprisingly well, exhibiting high
short-circuit photocurrents that exceed 20 mA cm−2 under full sunlight, the highest
certified PCE being 7.5%.
Multiple exciton generation (MEG) can be obtained from the absorption of a
single photon by a quantum dot if the photon energy is at least two times higher
than its bandgap (Nozik, 2003; Schaller and Klimov, 2004). The challenge is to
find ways to collect the excitons before they recombine. As recombination occurs
on a picosecond time scale, the use of mesoporous oxides as electron collectors
presents a promising strategy, because the electron transfer from the quantum dot
to the conduction band of the oxide electrode occurs within femtoseconds (Plass
et al., 2002). This opens up research avenues that ultimately may lead to photon
converters reaching external quantum efficiencies over 100%. Recent results on
depleted heterojunctions of PbSe(ZnO) confirm this expectation (Semonin et al.,
2011). A calculation based on Henry’s model (Henry, 1980) shows that the maxi-
mum conversion efficiency of a single-junction cell could be increased from 34%
to 44% by exploiting MEG effects.
where n is the ideality factor, whose value is between 1 and 2 for DSSCs, and
i o is the reverse saturation current. Thus for each order of magnitude decrease
in the dark current, the gain in V oc would be between 59 and 118 mV at room
temperature. Work in this direction is indispensable to raise the efficiency of
the DSSC significantly over the 15% limit with the currently employed redox
electrolytes.
Figure 9.12 Top: Cross-sectional SEM photograph of a mesoscopic solar cell using
CH3 NH3 PI3 as light harvester. The perovskite is infiltrated into the mesopores of the TiO2
scaffold. FTO and HTM signify fluorine-doped tin dioxide and hole-transport material.
Bottom: i–V curve of the best-performing device.
tests of these modules and polycrystalline silicon (pc-Si) have been running for
several years (Figure 9.13 shows a photograph of the test station). The test results
revealed the advantages of the DSSC as compared with silicon modules under
realistic outdoor conditions: for equal rating under standard test conditions (STC),
the DSSC modules produced 20–30% more energy than pc-Si modules (Toyoda,
2006). Recent data presented by Sony (Noda, 2012) show that today’s DSSC mod-
ule, having a standard AM 1.5 conversion efficiency of 10%, is expected to produce
on average over one year the same electricity as a 15% standard-rated pc Si module.
The superior performance of the DSSC can be ascribed to the following factors:
Figure 9.13 Outdoor field tests of DSSC modules by Fujikura Corporation near Tokyo,
Japan.
is more efficient than pc-Si in diffuse light or cloudy conditions. This advantage
is further enhanced for glass modules, which collect light in a bifacial way
from all angles.
While it is up to the commercial supplier to set the final price for such modules,
it is clear that the DSSC shares the cost advantages of all thin-film devices. In addi-
tion it uses no high-vacuum, cost-intensive steps in manufacture and only cheap
and readily available materials. Although it might be thought that the ruthenium-
based sensitiser adds high material cost, its contribution to cost is < US$ 0.01 W−1 p
because of the small amount used. Also, the present organic sensitisers match or
even surpass the performance of Ru complexes. Given these advantages at compa-
rable conversion efficiency, module costs below 0.5 W−1 p are realistic targets even
for production plants having well below GW capacity. The DSSC has thus become
a viable contender for large-scale solar energy conversion systems on the basis of
cost, efficiency, stability and availability, as well as environmental compatibility.
The British company G24i has started commercial sales of a flexible DSSC model.
Today the production rate is in the MW per year range.
9.7 Outlook
Using a principle derived from natural photosynthesis, mesoscopic injection solar
cells and in particular the DSSC, have become a credible alternative to solid-state
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch09 page 444
p-n junction devices. Conversion efficiencies of 12.3% and 15% have already been
obtained with single-junction liquid and solid-state mesoscopic solar cells, respec-
tively, on the laboratory scale, but there is ample room for further improvement.
Future research will focus on improving short-circuit current by extending the light
response of the sensitisers in the near-IR spectral region. Substantial gains in the
open-circuit voltage are expected from introducing ordered oxide mesostructures
and controlling the interfacial charge recombination by judicious engineering on
the molecular level. Hybrid cells based on solid-state inorganic and organic hole
conductors are an attractive option, in particular for the flexible DSSC embodiment.
Solar cells based on perovskite light absorbers are showing very promising effi-
ciencies and are rapidly developing into a particularly exciting new research area.
Mesoscopic dye-sensitised cells are well suited for a whole realm of appli-
cations, ranging from the low-power market to large-scale applications. Their
excellent performance in diffuse light gives them a competitive edge over silicon
in providing electric power for stand-alone electronic equipment both indoors and
outdoors. Dye-sensitized solar cells are already being applied in building-integrated
PV and this will become a fertile field of future commercial development.
As the epigraph at the chapter opening shows, at the 1912 IUPAC conference
in Washington a hundred years ago, the famous Italian photochemist Professor
Giacomo Ciamician predicted that mankind would unravel the secrets of photo-
synthesis and apply the principles used by plants to harvest solar energy in glass
buildings. His visionary thoughts now appear close to becoming a reality.
Acknowledgements
We gratefully acknowledge the insights and support from the many co-workers
with whom we have had the pleasure of working on these devices, and particularly
thank Dr Brian O’Regan for his help in preparing this manuscript.
References
3GSolar (2011), ‘3GSolar announces record outdoor reliability of third generation photo-
voltaic technology’, http:// www.3Gsolar.com.
Anderson N. A. and Lian T. Q. (2005), ‘Ultrafast electron transfer at the molecule-
semiconductor nanoparticle interface’, Annu. Rev. Phys. Chem. 56, 491–519.
Anderson S., Constable E. C., Dare-Edwards M. P., Goodenough J. B., Hamnett A. and
Seddon K. R. (1979), ‘Chemical modification of a titanium(IV) oxide electrode to give
stable dye sensitization without a supersensitiser’, Nature 280, 571–573.
Arakawa H., Yamaguchi T., Okada K., Matsui K., Kitamura T. and Tanabe N. (2009), ‘Highly
durable dye-sensitised solar cells’, Fujikura Tech. Rev. 2009, 55–59.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch09 page 445
Bach U., Lupo D., Comte P., Moser J. E., Weissortel F. and Salbeck J. (1998), ‘Solid-state
dye-sensitised mesoporous TiO2 solar cells with high photon-to-electron conversion
efficiencies’, Nature 395, 583–585.
Barbe C. J., Arendse F., Comte P., Jirousek M., Lenzmann F. and Shklover V. (1997),
‘Nanocrystalline titanium oxide electrodes for photovoltaic applications’, J. Amer.
Ceramic Soc. 80, 3157–3171.
Barkhouse D. A. R., Debnath R., Kramer, I. J., Zhitomirsky D., Pattantyus-Abraham A. G.,
Levina L., Etgar L., Grätzel M. and Sargent E. H. (2011), ‘Depleted bulk heterojunction
colloidal quantum dot photovoltaics’, Adv. Mater. 23, 3134–3140.
Bessho T., Zakeeruddin S. M., Yeh C.-Y., Diau E. W. G. and Grätzel M. (2010), ‘Highly
efficient mesoscopic dye-sensitised solar cells based on donor-acceptor-substituted
porphyrins’, Angew. Chem. Int. Ed. 49, 6646–6649.
Bisquert J. (2002), ‘Theory of the impedance of electron diffusion and recombination in a
thin layer’, J. Phys. Chem. B 106, 325–333.
Burnside S. D., Shklover V., Barbe C., Comte P., Arendse F. and Brooks K. (1998),
‘Self-organisation of TiO2 nanoparticles in thin films’, Chem. Mater. 10, 2419–2425.
Burschka A., Dualeh F., Kessler F., Baranoff E., Cevey-Ha N.-L., Yi Md. C., Nazeeruddin
M. K. and Grätzel M. (2011), ‘Tris(2-(1H-pyrazol-1-yl) pyridine) cobalt(III) as p-type
dopant for organic semiconductors and its application in highly efficient solid-state
dye-sensitised solar cells’, J. Amer. Chem. Soc. 133, 18042–18045.
Burschka J., Pellet N., Moon S. J., Humphry-Baker R., Gao P., Nazeeruddin M. K. and
Grätzel M. (2013), ‘Sequential deposition as a route to high-performance perovskite-
sensitized solar cells’, Nature 499, 316–319.
Casanova D., Rotzinger F. P. and Grätzel M. (2010), ‘Computational study of promising
organic dyes for high-performance sensitized solar cells’, J. Chem. Theory Comput. 6,
1219–1227.
Cass M. J., Qiu F. L., Walker A. B., Fisher A. C. and Peter L. M. (2003), ‘Influence of
grain morphology on electron transport in dye sensitised nanocrystalline solar cells’,
J. Phys. Chem. B 107, 113–119.
Chondroudis K. and Mitzi D. B. (1999), ‘Electroluminescence from an organic-inorganic
perovskite incorporating a quaterthiophene dye within lead halide perovskite layers’,
Chem. Mater. 11, 3028–3030.
Chung I., Lee B., He J., Chang R. P. H. and Kanatzidis M. G. (2012), ‘All-solid-state dye-
sensitized solar cells with high efficiency’, Nature 485, 486–489.
Crossland E. J. W., Nakita N., Sivaram V., Leijtens T., Alexander-Webber J. A. and Snaith
H. J. (2013), ‘Mesoporous TiO2 single crystals delivering enhanced mobility and opto-
electronic device performance’, Nature 495, 215–219.
Dare-Edwards M. P., Goodenough J. B., Hamnett A., Seddon K. R. and Wright R. D. (1980),
‘Sensitization of semiconducting electrodes with ruthenium-based dyes’, Faraday Dis-
cussions 70, 285–298.
Desilvestro J., Grätzel M., Kavan L., Moser J. and Augustynski J. (1985), ‘Highly efficient
sensitization of titanium dioxide’, J. Amer. Chem. Soc. 107, 2988–2990.
De Wild-Scholten M. J. and Veltkamp A. C. (2007), ‘Environmental life cycle analysis
of dye sensitised solar devices: status and outlook’, 22nd EU-PVSEC, https://www.
ecn.nl/publicaties/ECN-M-07-081.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch09 page 446
Dung D. H., Serpone N. and Grätzel M. (1984), ‘Integrated systems for water cleavage by
visible-light sensitization of TiO2 particles by surface derivatization with ruthenium
complexes’, Helvetica Chim. Acta 67, 1012–1018.
Durr M., Bamedi A., Yasuda A. and Nelles G. (2004), ‘Tandem dye-sensitised solar cell for
improved power conversion efficiencies’, Appl. Phys. Lett. 84, 3397–3399.
Durr M., Schmid A., Obermaier M., Rosselli S., Yasuda A. and Nelles G. (2005), ‘Low-
temperature fabrication of dye-sensitised solar cells by transfer of composite porous
layers’, Nature Materials 4, 607–611.
Durrant J. R., Haque S. A. and Palomares E. (2006), ‘Photochemical energy conversion:
from molecular dyads to solar cells’, Chem. Comm. 31, 3279–3289.
Dyesol (2013), ‘Dyesol exceeds stringent PV durability test by 400%’, May 28, www.
dyesol.com / posts / cat / corporate-news / post / dyesol-exceeds-stringent-pv-durability-
test-by-400-per-cent/.
Etgar L., Gao P., Xue Z., Peng Q., Chandiran A. K., Liu B., Nazeeruddin M. K. and Grätzel M.
(2012), ‘Mesoscopic CH3 NH3 PbI3 /TiO2 heterojunction solar cells’, J. Amer. Chem.
Soc. 134, 17396–17399.
Fabregat-Santiago F., Bisquert J., Garcia-Belmonte G., Boschloo G. and Hagfeldt A.
(2005), ‘Influence of electrolyte in transport and recombination in dye-sensitised
solar cells studied by impedance spectroscopy’, Solar Energy Mater. Solar Cells 87,
117–131.
Feldt S. M., Gibson E. A., Gabrielsson E., Sun L., Boschloo G. and Hagfeldt A. (2010),
‘Design of organic dyes and cobalt polypyridine redox mediators for high-efficiency
dye-sensitized solar cells’, J. Amer. Chem. Soc. 132, 16714–16724.
Ferrere S., Zaban A. and Gregg B. A. (1997), ‘Dye sensitization of nanocrystalline tin oxide
by perylene derivatives’, J. Phys. Chem. B 101, 4490–4493.
Frank A. J., Kopidakis N. and van de Lagemaat J. (2004), ‘Electrons in nanostructured TiO2
solar cells: transport, recombination and photovoltaic properties’, Coordination Chem.
Rev. 248, 1165–1179.
Gerischer H. and Tributsch H. (1968), ‘Elektrochemische Untersuchungen über den Mecha-
nismus der Sensibilisierung und Übersensibilisierung an ZnO-Einkristallen’, Ber. Bun-
senges. Phys. Chem. 72, 437–445.
Grätzel M. (2006), ‘Photovoltaic performance and long-term stability of dye-sensitised
mesosocopic solar cells’, Comptes Rend. Chimie 9, 578–583.
Grätzel M. (2009), ‘Recent advances in sensitised mesoscopic solar cells’, Acc. Chem. Res.
42, 1788–1798.
Grätzel M., Janssen R. A. J., Mitzi D. B. and Sargent E. H. (2012), ‘Materials interface
engineering for solution-processed photovoltaics’, Nature 488, 304–312.
Green A. N. M., Palomares E., Haque S. A., Kroon J. M. and Durrant J. R. (2005), ‘Charge
transport versus recombination in dye-sensitised solar cells employing nanocrystalline
TiO2 and SnO2 films’, J. Phys. Chem. B 109, 12525–12533.
Green M. A., Emery K., Hishikawa Y. and Warta W. (2011), ‘Solar efficiency tables (Ver-
sion 37)’, Progr. Photovoltaics 19, 84–92.
Hagfeldt A., Boschloo G., Sun L., Kloo L. and Pettersson H. (2010), ‘Dye-sensitized solar
cells’, Chem. Rev. 110, 6595–6663.
Hagfeldt A. and Grätzel M. (2000), ‘Molecular photovoltaics’, Acc. Chem. Res. 33,
269–277.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch09 page 447
Han L., Islam A., Chen H., Malapaka C., Chiranjeevi B., Zhang S., Yang X. and Yanagida M.
(2012), ‘High-efficiency dye-sensitised solar cell with a novel co-adsorbent’, Energy
Env. Science 5, 6057–6060.
Haque S. A., Palomares E., Cho B. M., Green A. N. M., Hirata N. and Klug D. R. (2005),
‘Charge separation versus recombination in dye-sensitised nanocrystalline solar cells:
the minimization of kinetic redundancy’, J. Amer. Chem. Soc. 127, 3456–3462.
Hara K., Sato T., Katoh R., Furube A., Yoshihara T. and Murai M. (2005), ‘Novel conjugated
organic dyes for efficient dye-sensitised solar cells’, Adv. Functional Mat. 15, 246–252.
Harikisun R. and Desilvestro H. (2011), ‘Long-term stability of dye solar cells’, Solar
Energy 85, 1179–1188.
Hengerer R., Kavan L., Krtil P. and Grätzel M. (2000), ‘Orientation dependence of
charge-transfer processes on TiO2 (anatase) single crystals’, J. Electrochem. Soc. 147,
1467–1472.
Henry C. H. (1980), ‘Limiting efficiencies of ideal single and multiple energy-gap terrestrial
solar cells’, J. Appl. Phys. 51, 4494–4500.
Hinsch A., Kroon J. M., Kern R., Uhlendorf I., Holzbock J. and Meyer A. (2001), ‘Long-term
stability of dye-sensitised solar cells’, Progr. Photovoltaics 9, 425–438.
Hirasawa M., Ishihara T., Goto T., Uchida K. and Miura N. (1994), ‘Magnetoabsorption
of the lower exciton in perovskite–type compound CH3 NH3 PbI3 ’, Physica B 291,
427–430.
Horiuchi T., Miura H., Sumioka K. and Uchida S. (2004), ‘High efficiency of dye-sensitised
solar cells based on metal-free indoline dyes’, J. Amer. Chem. Soc. 126, 12218–12219.
Hoyer P. and Weller H. (1995), ‘Potential-dependent electron injection in nanoporous col-
loidal ZnO films’, J. Phys. Chem. 99, 14096–14100.
Hsieh C.-P., Lu H.-P., Chiu C.-L., Lee C.-W., Chuang S.-H., Mai C.-L., Yen W.-N., Hsu S.-J.,
Diau E. W.-G. and Yeh C.-Y. (2010), ‘Synthesis and characterization of porphyrin sen-
sitizers with various electron-donating substituents for highly efficient dye-sensitized
solar cells’, J. Mater. Chem. 20, 1127–1134.
Huynh W. U., Dittmer J. J. and Alivisatos A. P. (2002), ‘Hybrid nanorod-polymer solar
cells’, Science 295, 2425–2427.
Im J.-H., Lee C. R., Lee J. W., Park S. W. and Park N. G. (2011), ‘6.5% efficient perovskite
quantum-dot-sensitized solar cell’, Nanoscale 3, 4088–4093.
Imahori H., Umeyama T. and Ito S. (2009), ‘Large π-aromatic molecules as potential sensi-
tizers for highly efficient dye-sensitized solar cells’, Acc. Chem. Res. 42, 1809–1818.
Ince M., Cardinali F., Raugossi M., Yum J., Gouloumis A., Martinez-Diaz M., Torre G.,
Nazeeruddin M., Grätzel M. and Torres T. (2012), ‘Phthalocyanines for molecular
photovoltaics’, Abstracts of Papers, 243rd ACS National Meeting and Exposition, San
Diego, CA, 25–29 March 2012, FUEL-522.
Ito S., Miura H., Uchida S., Takata M., Sumioka K., Liska P., Comte P., Pechy P. and
Grätzel M. (2008), ‘High conversion efficiency organic dye-sensitised solar cells with
a novel indoline dye’, Chem. Commun. 41, 5194–5196.
Ito S., Zakeeruddin S. M., Humphry-Baker R., Liska P., Charvet R. and Comte P. (2006),
‘High-efficiency organic-dye-sensitised solar cells controlled by nanocrystalline TiO2
electrode thickness’, Adv. Mater. 18, 1202–1205.
Jiu J. T., Wang F. M., Isoda S. and Adachi M. (2005), ‘Highly efficient dye-sensitised solar
cells based on single crystalline TiO2 nanorod film’, Chem. Lett. 34, 1506–1507.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch09 page 448
Pattantyus-Abraham A. G., Kramer I. J., Barkhouse A. R., Wang Y., Konstantatos G.,
Debnath R., Levina L., Raabe I., Nazeeruddin M. K., Grätzel M. and Sargent E.
H. (2010), ‘Depleted-heterojunction colloidal quantum dot solar cells’, ACS-Nano 6,
3374–3380.
Peter L. (2009) “‘Sticky electrons” transport and interfacial transfer of electrons in the
dye-sensitized solar cell’, Accounts Chem. Res. 42, 1839–1847.
Peter L. M. and Wijayantha K. G. U. (2000), ‘Electron transport and back reaction in dye-
sensitised nanocrystalline photovoltaic cells’, Electrochim. Acta 45, 4543–4551.
Plass R., Pelet S., Krueger J., Grätzel M. and Bach U. (2002), ‘Quantum dot sensitization
of organic–inorganic hybrid solar cells’, J. Phys. Chem. B 106, 7578–7580.
Qin H., Wenger S., Xu M., Gao F., Jing X., Wang P., Zakeeruddin S.-M. and Grätzel M.
(2008), ‘An organic sensitiser with a fused dithienothiophene unit for efficient and
stable dye-sensitised solar cells’, J. Amer. Chem. Soc. 130, 9202–9203.
Qin, Y. C. and Peng, Q. (2012), ‘Ruthenium sensitizers and their applications in dye
sensitized solar cells’, Int. J. Photoenergy, http://www.hindawi.com/journals/ijp/
2012/291579/.
Reddy P. Y., Giribabu L., Lyness C., Snaith H. J., Vijaykumar C., Chandrasekharam M.,
Lakshmikantam M., Yum J.-H., Kalyanasundaram K., Grätzel M. and Nazeerud-
din M. K. (2007), ‘Efficient sensitization of nanocrystalline TiO2 films by a near-
IR-absorbing unsymmetrical zinc phthalocyanine’, Angewandte Chem. Int. Ed. 46,
373–376.
Redmond G. and Fitzmaurice D. (1993), ‘Spectroscopic determination of flat-band poten-
tials for polycrystalline TiO2 electrodes in nonaqueous solvents’, J. Phys. Chem. 97,
1426–1430.
Rensmo H., Keis K., Lindstrom H., Sodergren S., Solbrand A. and Hagfeldt A. (1997),
‘High light-to-energy conversion efficiencies for solar cells based on nanostructured
ZnO electrodes’, J. Phys. Chem. B 101, 2598–2601.
Rhee Y. M. and Head-Gordon M. (2007), ‘Scaled second-order perturbation correction
to configuration interaction singles: efficient and reliable excitation energy methods’,
J. Phys. Chem. 111, 5314–5326.
Rothenberger G., Fitzmaurice D. and Grätzel M. (1992), ‘Optical electrochemistry. 3.
Spectroscopy of conduction-band electrons in transparent metal-oxide semiconductor
films — optical determination of the flat-band potential of colloidal titanium dioxide
films’, J. Phys. Chem. 96, 5983–5986.
Sargent E. H. (2012), ‘Colloidal quantum dot solar cells’, Nat. Photonics 6, 133–135.
Sauvage F., Chhor S., Marchioro A., Moser J.-E. and Grätzel M. (2011), ‘Butyronitrile-based
electrolyte for dye-sensitised solar cells’, J. Amer. Chem. Soc. 133, 13103–13109.
Sayama K., Sugihara H. and Arakawa H. (1998), ‘Photoelectrochemical properties of
a porous Nb2 O5 electrode sensitised by a ruthenium dye’, Chem. Mater. 10,
3825–3832.
Schaller R. D. and Klimov V. I. (2004), ‘High efficiency carrier multiplication in
PbSe nanocrystals: implications for solar energy conversion’, Phys. Rev. Lett. 92,
86601–86604.
Schmidt-Mende L., Bach U., Humphry-Baker R., Horiuchi T., Miura H. and Ito S. (2005a),
‘Organic dye for highly efficient solid-state dye-sensitised solar cells’, Adv. Materials
17, 813–815.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch09 page 451
Schmidt-Mende L., Kroeze J. E., Durrant J. R., Nazeeruddin M. K. and Grätzel M. (2005b),
‘Effect of hydrocarbon chain length of amphiphilic ruthenium dyes on solid-state dye-
sensitised photovoltaics’, Nano Lett. 5, 1315–1320.
Semonin O. E., Luther J. M., Choi S., Chen H.-Y., Gao J., Nozik A. J. and Beard M. C.
(2011), ‘Peak external quantum efficiency exceeding 100% via MEG in a quantum dot
solar cell’, Science 334, 1530–1533.
Schmickler W. (1996), Interfacial Electrochemistry, Oxford University Press, Oxford.
Shklover V., Ovchinnikov Y. E., Braginsky L. S., Zakeeruddin S. M. and Grätzel M. (1998),
‘Structure of organic/inorganic interface in assembled materials comprising molecu-
lar components. Crystal structure of the sensitiser bis[(4,4 -carboxy-2,2 -bipyridine)
(thiocyanato)]ruthenium(II)’, Chem. Mater. 10, 2533–2541.
Snaith H. J., Moule A. J., Klein C., Meerholz K., Friend R. H. and Grätzel M. (2007), ‘Effi-
ciency enhancements in solid-state hybrid solar cells via reduced charge recombination
and increased light capture’, Nano Letters 7, 3372–3376.
Snaith H. J., Zakeeruddin S. M., Schmidt-Mende L., Klein C. and Grätzel M. (2005), ‘Ion-
coordinating sensitiser in solid-state hybrid solar cells’, Angew. Chem. Int. Ed. 44,
6413–6417.
Stathatos E., Lianos R., Zakeeruddin S. M., Liska P. and Grätzel M. (2003), ‘A quasi-solid-
state dye-sensitised solar cell based on a sol–gel nanocomposite electrolyte containing
ionic liquid’, Chem. Mater. 15, 1825–1829.
Tennakone K., Kumara G. R. R. A., Kumarasinghe A. R., Wijayantha K. G. U. and
Sirimanne P. M. (1995), ‘A dye-sensitised nanoporous solid-state photovoltaic cell’,
Semiconductor Sci. Technol. 10, 1689–1693.
Toyoda T. (2006), ‘Outdoor tests of large area dye-sensitised solar cell modules by Aisin
Seiki, Inc.’, Re2009 Renewable Energy Congress, Makura, Japan, October.
Tsubomura H., Matsumura M., Nomura Y. and Amamiya T. (1976), ‘Dye-sensitised zinc
oxide–aqueous electrolyte–platinum photocell’, Nature 261, 402–403.
Vittadini A., Selloni A., Rotzinger F. P. and Grätzel M. (1998), ‘Structure and energetics
of water adsorbed at TiO2 anatase (101) and (001) surfaces’, Phys. Rev. Lett. 81,
2954–2957.
Wang P., Klein C., Humphry-Baker R., Zakeeruddin S. M. and Grätzel M. (2005a), ‘A high
molar extinction coefficient sensitiser for stable dye-sensitised solar cells’, J. Amer.
Chem. Soc. 127, 808–809.
Wang P., Klein C., Humphry-Baker R., Zakeeruddin S. M. and Grätzel M. (2005b), ‘Stable
8% efficient nanocrystalline dye-sensitised solar cell based on an electrolyte of low
volatility’, Appl. Phys. Lett. 86, 1235081–1235083.
Wang P., Zakeeruddin S. M., Humphry-Baker R., Moser J. E. and Grätzel M. (2003a),
‘Molecular-scale interface engineering of TiO2 nanocrystals: improving the efficiency
and stability of dye-sensitised solar cells’, Adv. Mater. 15, 2101–2103.
Wang P., Zakeeruddin S. M., Moser J. E., Nazeeruddin M. K., Sekiguchi T. and Grätzel M.
(2003b), ‘A stable quasi-solid-state dye-sensitised solar cell with an amphiphilic ruthe-
nium sensitiser and polymer gel electrolyte’, Nat. Materials 2, 402–407.
Walter M. G., Rudine A. B. and Wamser C. C. (2010), ‘Porphyrins and phthalocyanines in
solar photovoltaic cells’, J. Porphyrins and Phthalocyanines 14, 759–792.
Wells H. L. (1893), ‘Uber die Casium- und Kalium-Blei halogenide’, Z. Anorg. Chem. 3,
195–210.
September 17, 2014 11:3 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch09 page 452
West W. (1874), Vogel Centennial Symp., Photogr. Sci. Eng. 18, 35.
Wenger S., Seyrling S., Tiwari A. N. and Graetzel M. (2009), Fabrication and performance
of a monolithic dye-sensitized TiO2 /Cu(In,Ga)Se2 thin film tandem solar cell, Appl.
Phys. Lett. 94, 173508/1–173508/3.
Yella A., Lee H.-W., Tsao H. N, Yi C., Chandiran A. K., Nazeeruddin M. K., Diau E. W.-G.,
Yeh C.-Y., Zakeeruddin S. M. and Grätzel M. (2011), ‘Porphyrin-sensitised solar cells
with cobalt (II/III)-based redox electrolyte exceed 12% efficiency’, Science 334, 629–
634.
Yum J.-H., Baranoff E., Kessler F., Moehl T., Ahmad S., Bessho T., Marchioro A., Ghadiri E.,
Moser J.-E., Yi C., Nazeeruddin M. K. and Grätzel M. (2012), ‘A cobalt complex redox
shuttle for dye-sensitised solar cells with high open circuit potentials’, Nature Commun.
3, Art. 631.
Yum J.-H., Hagberg D. P., Moon S.-J., Karlsson K. M., Marinado T., Sun L., Hagfeldt A.,
Nazeeruddin M. K. and Grätzel M. (2009), ‘A light-resistant organic sensitiser for
solar-cell applications’, Angew. Chem. Int. Ed. 48, 1576–1580.
Zaban A., Ferrere S. and Gregg B. A. (1998), ‘Relative energetics at the semiconductor
sensitizing dye electrolyte interface’, J. Phys. Chem. B 102, 452–460.
Zhang G., Bala H., Cheng Y., Shi D., Lv X., Yu Q. and Wang P. (2009), ‘High efficiency
and stable dye-sensitised solar cells with an organic chromophore featuring a binary
π-conjugated spacer’, Chem. Comm. 16, 2198–2200.
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch10 page 453
CHAPTER 10
10.1 Introduction
The quantum well solar cell (QWSC) was proposed by Barnham and Duggan
(1990) as a new type of multiple-bandgap, single-junction solar cell. The principle
is similar to that of the monolithic tandem cell: several bandgaps are used to absorb
different spectral ranges efficiently. However, rather than using two junctions made
from different semiconductors, the QWSC uses ultra-thin layers of different mate-
rials in a monolithic, two-terminal arrangement. The technology of quantum wells
(QWs) is borrowed from the optoelectronics industry. The 1980s and 1990s saw
developments in the use of low-dimensional semiconductor structures in optical
modulators, photodetectors, lasers and other devices (Weisbuch and Vinter 1991;
Zory, 1993). These applications exploit the special optical and electronic properties
available with highly confined carrier populations. Before 1989, low-dimensional
systems had been considered in connection with solar cells only in the context of
improved carrier transport through the neutral regions. In the QWSC, QWs are used
in the active region of the solar cell to enhance solar photon absorption and boost
the photocurrent. The QWs can be considered approximately as an incremental
current source in parallel with a conventional homojunction solar cell.
The use of low-dimensional semiconductor structures in solar cells was first
proposed in the early 1980s, where QW and superlattice arrays were introduced
into the doped regions of a tandem device based on p-n junctions (Wanlass and
Blakeslee 1982; Chaffin et al., 1984; Goradia et al., 1985). However, poor collection
due to inefficient carrier diffusion within a doped QW structure led to performances
inferior to those of bulk alloys, and the idea was abandoned.
453
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch10 page 454
The usual and most effective geometry is to insert an array of QWs called
a multiple QW (MQW) into the depletion region (rather than the doped region)
of a p-n- or p-i -n-junction solar cell. Quantum wells absorb more solar photons
and so enhance the photocurrent. However, because of their lower bandgap, QWs
also enhance the dark recombination current that opposes the photocurrent. If
the increase in photocurrent exceeds the increase in dark current under operating
conditions, then the power-conversion efficiency of the cell will be increased.
The possibility of a higher limiting efficiency than is possible with a simple
single-junction cell is the main reason for interest in the QWSC, but QW structures
have also been considered as a means of improving the performance of practical
solar cells. For instance, the ability to control the bandgap through QW width is
useful in various applications, and the response of QW cells to increases in temper-
ature makes them attractive for solar concentration. The properties of QWSCs have
been studied, theoretically and experimentally, by a number of research groups.
Significant milestones in the development of the QWSC include the demonstration
of increased photocurrent from photogeneration in the quantum wells of a QWSC
(Barnham et al., 1991; Freundlich et al., 1994) and an understanding of the carrier
escape and current transport processes in forward bias (Nelson et al., 1993; Kitatani
et al., 1995). In parallel, the opportunity for a fundamental efficiency advantage was
debated (Corkish and Green, 1993; Araujo et al., 1994; Ragay et al., 1994; Luque
et al., 2001) and full device models were developed (Renaud et al., 1994; Anderson,
1995; Nelson, 1995; Ramey and Khoie, 2003). More recently, developments in so-
called ‘strain-balanced’ QW structures enabled high efficiency single-junction and
tandem devices. This development made QWSC technology potentially viable for
applications in space and terrestrial concentrator systems. Quantum well solar cell
technology is now being commercialised by JDS Uniphase Corporation. Progress
in epitaxial growth of highly strained layers has enabled strain-balanced superlat-
tice solar cells to be achieved (Wanget al., 2012).
This chapter covers the following areas: QW technology and materials issues;
the physics of QWs and how they differ from bulk materials in the processes
important for solar energy conversion; experimental results showing how these
properties affect the performance of the solar cell; the question of the limits to
efficiency; and practical applications of QWSCs.
p
n
n
EF
p qV Eb
EF
Figure 10.1 Schematic diagram of energy band vs. distance for a p-i-n QWSC at forward
bias V . Quantum wells are placed in the undoped i-layer, where the electric field helps to
separate photogenerated electron–hole pairs. Under illumination, the charges are separated
p
by the electric field and the electron and hole quasi-Fermi levels E F and E Fn are split by qV,
as shown.
to extend the field-bearing region. The rationale is that QWs should be placed in
this region where charge separation and collection is most efficient.
The QWs are thin layers of a second, narrower gap semiconductor between
barrier layers of the host material. III–V semiconductors are normally used for
both QW and host material. Typically the QWs are 60–150 Å wide, separated by
barriers of 50 Å or more. Some 50 QWs can be placed in an i -region 0.5–1 µm
thick. In practice the i -layer thickness is limited by the background level of charged
impurities, since these can cause the electric field to fall to zero within the i -layer
and render the remaining, neutral part of the junction useless (Zachariou et al.,
1996; Barnham et al., 1991).
The remaining parts of the cell — the p-layer, n-layer, substrate, window lay-
ers, contacts and optical coatings — are modelled on conventional III–V solar cell
designs. The emitter (top layer) may be somewhat narrower in order to admit as
much light as possible to the active i -region. Otherwise, the same design considera-
tions apply: choice of polarity ( p-n or n- p); techniques to enhance minority-carrier
collection such as surface passivation, window layers and graded emitters; choice
of materials for dopant, antireflective (AR) coat and substrate. The original goal
was to use the incremental photocurrent provided by the QWs to enhance the
efficiency of a well-designed homojunction cell.
An additional advantage is the possibility of controlling the QW width to tune
the absorption edge of one or more cells in a multijunction device.
10.2.2 Materials
Quantum well solar cells have been studied in several III–V materials combi-
nations. A short summary of some key examples follows, expressed as Barrier/
QW material: AlGaAs/GaAs (Barnham et al., 1991), GaAs/InGaAs (Barnes
et al., 1996), InP/InGaAs (Zachariou et al., 1998), InGaP/GaAs (Barnham et al.,
1996), InP/InAsP (Monier et al., 1998), InGaAsP/InGaAs (Rohr et al., 2006) and
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch10 page 456
Figure 10.2 Plot of bandgap vs. lattice constant for III–V semiconductors.
of elements has been observed, this is considered not to be a major limitation for
a functioning device (Afshar et al., 2011).
Si-based QWs and superlattices have also been investigated in an attempt to
fabricate an all-silicon tandem solar cell (Green, 2003), exploiting the 1.1 eV bulk
silicon junction as a bottom cell. Three-dimensional confinement is required to
achieve a 2 eV gap in Si for the top junction. For this purpose, arrays of Si QDs
confined by SiO and SiN barriers have been fabricated with an absorption threshold
approaching this value, although in common with II–VI QD structures, efficient
current transport through the QD array remains a challenging problem (Conibeer
et al., 2010).
energy
conduction band
Eb
Eg
valence band
well material
barrier material
distance
where k is the total wave vector, k is the component in the xy plane (such that
k2 = k2 + kn2 ), and m ∗ is the effective mass of the carrier in this plane.
In the envelope function approximation, the shift V (z) in the conduction or
valence band edge due to the QW is considered as a perturbation to the periodic
crystal potential, and the wave functions as crystal eigenfunctions modulated by
an ‘envelope function’. The confined state energies E n and envelope functions
Fn (z) are solutions to an ‘effective mass’ equation, which resembles Schrödinger’s
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch10 page 461
equation for a one-dimensional potential well. They are analogous to the energy
levels and wave functions of a one-dimensional quantum system. For a QW of
width L and depth V
h̄ 2 d 2 Fn (z)
− + E(z)Fn (z) = E n Fn (z)
2m ∗ dz 2
where
E(z) = 0, −L/2 ≤ z ≤ L/2
(10.2)
E(z) = V, z > |L/2|
This equation holds for both electrons in the conduction band and holes in the
valence band, but with different values of m ∗ and V . Energies E n are measured up
from the bottom of the QW in the conduction band for electrons, and down from
the top of the valence band for holes. The well depth V depends on the composition
of the barrier and well materials and on how the difference in bandgap is divided
between the valence and conduction bands. The effective mass m ∗ for each carrier
type is in general different for well and barrier.
In III–V semiconductors two different types of hole, heavy and light, need to
be considered. In the bulk crystal, heavy and light holes are carriers with different
effective mass associated with two degenerate crystal bands. For a QW in unstrained
material, heavy and light holes occupy the same potential well in the valence
band, but with different sets of confined-state energies on account of their different
effective masses. In a strained QW, the well depths for heavy and light holes can
be different. The number N of confined states contained in the QW for each carrier
type is given by
√
L 2m ∗ V
N = int +1 (10.3)
π h̄
where int(x) means the integer part of x. N increases with increasing well width and
depth, and carrier effective mass. The well is normally narrow enough to admit only
a few confined states. At energies E > V the carriers are no longer confined and a
continuum of states becomes available, as in the bulk material. These continuum
states will not be considered here.
In accordance with the uncertainty principle, the lowest energy level is always
shifted away from the bottom of the well, by an amount that increases with increas-
ing quantum confinement. This means that the ground-state energy, and hence the
absorption edge, can be controlled simply by varying the well width.
The corresponding envelope functions have well-defined parity and penetrate
further into the barrier as energy is increased. Energy levels and envelope functions
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch10 page 462
0.2 e3
e2
0 e1
-1 -0.5 0 0.5 1
-0.2
-0.4
Energy/eV
-0.6
-0.8
-1
-1.2
-1.4
hh2 hh1
hh3
hh4
-1.6
Well widths
Figure 10.4 Calculated energy levels and envelope functions for a 100 Å GaAs QW in
Al0.3Ga0.7As. The relative energies of confined states and bandgaps are to scale, and the
bottom of the conduction band is taken as the zero of energy. Quantum number is measured
up from the bottom of the well for electrons, and down from the top of the well for holes.
for a typical Alx Ga1−x As /GaAs QW are shown in Fig. 10.4. In the SL configura-
tion, neighbouring wells are coupled, extended-state envelope functions span the
entire SL, and the previously discrete energy levels of the QW broaden into bands.
These effects improve carrier transport in the growth direction.
The density of states function can be constructed from the energy spectrum
in the usual way. For a QW of width L the density of states per unit volume V is
given by
2 m∗
N
D(E) = δ[E − E(k)] = (E − E n ) (10.4)
V
k
π h̄ 2 L n=1
where δ is the Dirac delta function and is the Heaviside function. As illustrated in
Fig. 10.5, D(E) has the staircase structure characteristic of quasi-two-dimensional
systems.
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch10 page 463
D(E)
infinite well
E1 E2 E3 E
Figure 10.5 Schematic density of states function D(E) for a finite QW, compared with
that for an infinitely deep QW and for the well material in the bulk. The first three confined
state energies, E 1 , E 2 and E 3, are shown.
This allows us to calculate the concentrations n (of electrons) and p (of holes)
in the QW, assuming a local quasi-thermal equilibrium. For electrons with density
of states function Dcb (E) in the conduction band
∞
n= Dcb (E(k)) f F D (E(k), T,
µe )d E (10.5)
E cb
n = n iq exp[(
µe − E i + θn )/kT ] (10.6)
electric field
q L
L
Figure 10.6 Band profile for a QW subject to an electric field in the growth direction. As
the field is increased the right-hand barrier is reduced, increasing the probability of electron
escape by thermionic emission or tunnelling.
10.3.3 Photogeneration
In a solar cell, photon absorption across the bandgap is important. Fermi’s golden
rule gives the absorption coefficient α in terms of the confined-state energies and
overlap integrals. For transitions between an initial valence-band state |i , of energy
E i and a final, conduction-band state | f , of energy E f , under the influence of an
electromagnetic field of angular frequency ω and polarisation e we have (Bastard,
1988)
A
α(E) = | f |e.p|i |2 δ[E f − E i − E]( f F D (E i ) − f F D (E f )) (10.7)
ω
i, f
where E is the photon energy h̄ω, p is the quantum mechanical momentum operator
and A is a sample-dependent optical constant. In the usual case where the light is
incident normal to the plane of the QW, the matrix element is proportional to the
overlap integral Mlm between the initial valence sub-band, l, and final conduction
band, m, envelope functions
Mlm = Fel (z)Fhm (z)dz (10.8)
This means that optical transitions are allowed only between sub-bands of the
same parity (l and m both even or both odd), and are strong only when l = m.
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch10 page 465
In addition, coulombic bound states (excitons) are formed at an energy just below
the minimum for each optically allowed sub-band-to-sub-band transition. The exci-
tons appear as strong peaks in the spectrum, even at room temperature, because of
their higher binding energy in two-dimensional systems. Including only the prin-
cipal (1s) exciton and summing Eq. (10.7) over initial and final state energies for
the lth electron – mth hole sub-band pair, we have
αlm (E) = αlh/ hh Mlm [ flm δ(E − Elm − Blm ) + (E − E lm )] (10.9)
where Elm is the electron-hole transition energy before Coulombic effects are
included, Blm and f lm are the exciton binding energy and oscillator strength, respec-
tively, and the constants αlh/ hh represent the absorption coefficient on the first step
edge. In III–V semiconductors, optical transitions occur between both electron–
heavy hole (hh) and electron–light hole (lh) states. The total absorption is the sum
of contributions from all such transitions.
α(E) = αel hh m (E) + αel lh m (E) (10.10)
l,m l,m
where each electron–hole sub-band pair contributes a step function and a set of
excitons to the total absorption spectrum. The absorption coefficient for a typical
Alx Ga1−x As /GaAs QW is shown in Fig. 10.7.
The QW absorption spectrum thus reflects the step-like form of the density
of states, modified by strong excitonic peaks. (Because of the strong exciton, the
QW spectrum may have a steeper absorption edge than the equivalent bulk alloy,
6
GaAs
5
–1
Al0.33Ga 0.67As
4
0
1.3 1.4 1.5 1.6 1.7 1.8 1.9 2
Photon energy/eV
Figure 10.7 Calculated absorption coefficient for a 100 Å Al0.33 Ga0.67 As/GaAs QW com-
pared with the absorption of bulk GaAs and bulk Al0.33 Ga0.67 As. (For the QW, the absorp-
tion coefficient is per unit thickness of well material, not including barrier thickness.)
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch10 page 466
which could be useful for certain PV applications.) The absorption edge or effective
bandgap E a is blue-shifted from the absorption edge E g of the well material in the
bulk by the joint confinement energies E 11h of the lowest electron and heavy hole
sub-bands less the corresponding exciton binding energy B11h .
10.3.5 Recombination
The processes that govern recombination in bulk materials apply to QWs. For
III–Vs the most important, in practice, is non-radiative recombination through
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch10 page 467
traps. For a single trap state in the bandgap, the Shockley–Read–Hall recombination
rate is given by
np − n 2i
rnr = (10.12)
τe ( p + pt ) + τh (n + n t )
where pt , n t are the equilibrium populations of trap states occupied by holes and
electrons, and τ p , τn are the respective carrier trapping times. This formulation
should be appropriate to a QW provided that n and p are defined using the quasi-
Fermi level of the carriers in the QW (Eq. (10.6)). The lifetime parameters are
properties of the material and so as a first approximation to a QW we may take
the same values as for the well material in the bulk. However, the accumulation of
defects at the QW interface may affect the location and density of trap states, and
quantum confinement may reduce the trapping times.
In the limit of ideal material, radiative recombination is the process that deter-
mines solar cell efficiency. The excess radiative recombination in the biased device
(i.e. in addition to the recombination that balances thermal generation in equilib-
rium) then constitutes the dark current. In any volume element δV the radiative
recombination rate rrad depends on the local absorption spectrum α(E) and the
local quasi-Fermi level separation µ̃F , according to
rrad δV = α(E) j (E, T, µ F )d EδV (10.13)
where the emitted flux density j is given by the generalised Planck equation (Wür-
fel, 1982; Tiedje et al., 1984)
2nr2 E2
j (E, T, µ̃ F ) = (E− µ̃
(10.14)
3 2
h c e F )/ kT − 1
and n r is the local refractive index, h is Planck’s constant and c the speed of light.
apply the drift–diffusion equations to the composite structure and represent the
QW by a region of increased carrier population. In this approach it is assumed that
the QW and host are in quasi-thermal equilibrium and that the quasi-Fermi level
is continuous across the interfaces of the QW.
However, this treatment neglects the fact that carriers with energies below the
barrier band edge are essentially trapped in the QW. These carriers may escape
from the QW by quantum mechanical tunnelling or thermally assisted tunnelling
when the QW is subject to an electric field in the growth direction, and other car-
riers may be captured into the QW by scattering from higher energy states. Both
of these processes are quantum mechanical and cannot effectively be described
with the drift–diffusion equations. An alternative approach is to consider two pop-
ulations for each carrier: a mobile population with energy greater than the barrier
band edge, and a confined population in the QW. The confined population does not
actually contribute to the current, but it affects the current by influencing the rate
at which carriers are added to the mobile population through escape, or removed
from it through capture. Equation (10.15) for i e should therefore include terms for
escape and capture:
where Tr(k z ) is the probability of transmission through the barrier at wave vector
k z . The integral should be carried out with respect to wave vector rather than energy,
because of the anisotropy of the density of states. For shallow QWs in short-circuit
and low-forward-bias conditions, the escape current appears to be equivalent to the
sheet-generation rate in the QW (Nelson et al., 1993). In the operating conditions
of a QW, when barrier states are also populated, Eq. (10.17) should be modified to
allow for a finite probability of population of the final states, and a term added for
capture of carriers into the QW from states above the barrier. This is a notoriously
difficult problem, involving both localised and continuum states, and has not at the
time of writing been resolved for the QWSC.
Once given the expressions for carrier-pair generation, recombination and
current, carrier continuity may be invoked to complete the set of equations for the
QW device. Discretising the continuity equations for electrons in a QW of width L
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch10 page 469
where r is the front-surface reflectivity and ηesc the probability of carrier escape
from the QW. The fraction (1 − ηesc ) of the carriers that do not escape can be
assumed to recombine in the QW. In fact, for QWs in any of the systems mentioned
at room temperature, ηesc is unity. This has been confirmed by experimental studies
of the dependence on the field and temperature of the photocurrent from single
QW devices (Nelson et al., 1993; Barnes, 1994; Zachariou et al., 1998), which
show that carriers escape from the QWs at room temperature by thermally assisted
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch10 page 470
0.50
0.30
0.20
0.10
0.00
400 500 600 700 800 900
Wavelength/nm
Figure 10.8 Measured spectral response (external quantum efficiency) for an Al0.33 Ga0.67
As p-i-n test device with and without 50 87 Å QWs in the i-region. These devices were not
AR coated.
where SR(E) is the incremental SR due to the QWs, which is given by Eq. (10.19)
for E < E b and the difference between the SR of well and host material for E > E b .
To maximise the effect of the QWs on photocurrent we need an SR which is
as high as possible for E > E a . In principle this can be achieved by increasing the
number of QWs and reducing the optical depth of the top layer (the p-region in a
p-i -n structure) by reducing its thickness or compositional grading, as shown in
Fig. 10.9.
In practice it is not always straightforward to increase the number of QWs.
A wide i -region requires a relatively low level of charged impurities (<1015 cm−3 )
in order to maintain the electric field (Barnham et al., 1991; Zachariou et al., 1996).
Figure 10.10 shows how the QW SR is affected by high background doping, in
this case arising from diffusion of the zinc dopant into the i -region.
For GaAs/In x Ga1−x As QWs, strain limits the spacing of the QWs. The GaAs
barrier layers need to exceed a critical thickness several times the Inx Ga1−x As
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch10 page 471
1.00
0.80
Spectral response
0.00
400 500 600 700 800 900
Wavelength/nm
Figure 10.9 Calculated spectral response SR for an Al0.33 Ga0.67 As p-i-n device and a
series of Al0.33 Ga0.67 As/GaAs QW devices, showing the effect of: a) reducing the p-layer
thickness and increasing the i-layer thickness; b) adding an AR coat; c) adding a front-
surface window and back-surface reflector. The measured SR for a cell of design (b) is
compared with the calculation.
1.00
Model: High background doping
Model: Low background doping
0.80 Experimental data
Spectral response
0.60
0.40
0.20
0.00
400 600 800 1000 1200 1400 1600
Wavelength/nm
Figure 10.10 Measured and calculated SR for an InP p-i-n device containing 20
In0.53 Ga0.47 As QWs. The calculated curves show that an unintentional background doping
in the i−layer of around 2 × 1016 cm−3 is required to explain the low SR. Diffusion of zinc
dopants increases the space charge in the i-region so that at zero bias the electric field falls
to zero near the centre of the i-region and only around half of the carriers generated in the
QWs are collected.
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch10 page 472
Figure 10.11 Measured SR for a series of GaAs/Inx Ga1−x As devices, each with 10 QW,
processed in different ways. The low SR with no light trapping is substantially increased by
adding a plane or a rough gold mirror at the back surface. The rough mirror, a gold epoxy
resin, achieves light trapping effectively through large-angle internal scattering.
i Dk (V ) = i nr (V ) + i rad (V ) (10.21)
Since the generation rate g = 0 in the dark, i Dk is found by integrating all the
contributions to the recombination rate r over the volume of the cell. Radiative
recombination within the interior of the device is partly cancelled by absorption,
and the radiative current density i rad is given by q times the net photon flux escaping
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch10 page 473
through the device surface. Thus i rad is obtained by integrating Eq. (10.13) over
surface elements S and solid angle
∞ a
i rad = q (E, θ, s) j (E, T,
µ F )d
.dS.dE
0 S
10000
1000
100
Dark current / A m -2
85A MQW
10 50A SQW
AlGaAs p-i-n
1
140A SQW
0.4 0.6 0.8 1 1.2 1.4
0.1 GaAs p-i-n
GaAs / InGaAs
0.01
0.001
0.0001
Bias / V
Figure 10.12 Measured dark current densities for a series of QW and homogenous p-i-n
devices. The curves show that adding GaAs QWs to an Al x Ga1−x As p-i-n cell increases
the dark current, as does adding Inx Ga1−x As QWs to a GaAs device. The effect on the dark
current is larger for wider QWs, and for a greater number of QWs.
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch10 page 475
is maximal away from the centre of the i -region, and so reduce the recombination
rate at a centred QW. This would also explain the change in ideality factor, as
shown in Nelson et al. (1995, 1999). But it would not explain why the dark current
is lower than expected in multi-quantum well devices, where we would expect the
regular MQW array to sample all n/ p ratios equally.
A possible explanation is that n and p are smaller than expected in the QW,
i.e. the quasi-Fermi levels are closer to the centre of the bandgap than in the
neighbouring host material. This is supported by complementary measurements
of the radiative recombination current in single QW devices (Nelson et al., 1997,
1998), which show that the emission spectrum from a single QW is some tens of
meV smaller (and the radiative current a factor of 2–4 smaller) than expected if the
quasi-Fermi level separation were constant across the well/barrier interface.
This explanation suggests that the QWs are not in quasi-thermal equilibrium
with the barrier material. Quasi-equilibrium is appropriate when current is dom-
inated by low-field carrier drift and diffusion. These conditions are usually satis-
fied for a homojunction cell, where variations in carrier population are smooth.
In QWSCs in the region of the QW interfaces, as discussed above, current may
be dominated by carrier escape from the QWs through thermionic emission and
thermally assisted tunnelling. These currents may greatly exceed drift–diffusion
currents in the direction of increasing kinetic energy, and therefore the notion of
quasi-thermal equilibrium may not be valid here.
This explanation was proposed by Corkish and Honsberg (1997). They draw
on studies of bulk heterojunctions which show that a step in the minority-carrier
Fermi level may result at a heteroface in conditions where transport is limited by
thermionic emission or transport across a space-charge region. They show that a
moderately high level of background doping in the QW could give rise to reduced
quasi-Fermi level separation and lower dark currents. This is a promising idea and
more detailed modelling incorporating carrier escape and capture may well provide
a quantitative explanation of the observed dark currents.
Optimisation of the QWSC clearly requires minimisation of the incremental
recombination current. One way to do this would be to choose a material for the
QW with long non-radiative recombination times. In GaAs, for example, minority
carrier lifetimes are long compared to those in Alx Ga1−x As, and it is possible
that the recombination current from an Alx Ga1−x As /GaAs MQW may be lower
than that from the Alx Ga1−x As alloy of equivalent effective bandgap. Also the
dependence of recombination rate on QW position means that it should be possible
to optimise the design of QWSCs by locating QWs away from the centre of the
SCR, in the regions where n p or p n. (Nelson et al., 1999). This effect results
from the asymmetry in carrier populations and the form of the Shockley–Read–Hall
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch10 page 476
recombination rate; it does not require that the quasi-Fermi level in the QW be
reduced.
i (V ) = i sc − i Dk (V ) (10.25)
i (V ) = i sc + i sc − i Dk (V ) − i Dk (V ) (10.26)
Therefore, QWs can benefit the power conversion efficiency only when the
increase in photocurrent i sc exceeds the increase in recombination current i Dk
100000
Model: low background doping
Model: high background doping
Experimental data
1000
Dark current / A m-2
10
0.1
0.001
0.4 0.8 1.2 1.6
Bias / V
Figure 10.13 Measured and modelled dark currents for an Al0.34 Ga0.66 As/GaAs single
QW device. The black line shows the dark current expected when the background doping
level Ni in the i region is small (∼1014 cm−3 ) and the QW is located close to the plane where
electron and hole densities are equal. For high background doping (Ni ≈ 2 × 1016 cm−3 ),
p n at the QW and non-radiative recombination is reduced. The lower ideality factor
reflects the asymmetry in the electron and hole densities.
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch10 page 477
i sc − i Dk (V ) > 0 (10.27)
The advantage increases with the number of QWs since, while i sc increases
approximately linearly with the number of QWs, the decrease in open-circuit volt-
age due to i Dk changes only logarithmically.
Figures 10.14 and 10.15 show I –V characteristics for an Al0.3 Ga0.7 As p-i -n
cell with and without 30 GaAs QWs and for a GaAs p-i -n cell with and without 10
60
AlGaAs p-i-n cell
50
AlGaAs/ GaAs QW cell
Current density/A m–2
40
30
20
10
-10
-20
0 0.2 0.4 0.6 0.8 1 1.2
Bias/V
Figure 10.14 Measured current–voltage characteristics for an Al0.3 Ga0.7 As p-i-n device
with and without 30 GaAs QWs, in a white light source approximating to a 3200 K black-
body spectrum. Note that these devices were not AR-coated, hence the low current densities.
600
500
–2
Current density/A m
400
300
200
GaAs p-i-n cell
100 GaAs/InGaAs QW cell
-100
0 0.2 0.4 0.6 0.8 1 1.2
Bias/V
Figure 10.15 Measured current–voltage characteristics for a GaAs p-i-n cell with and
without 10 InGaAs QWs, using the same light source as for Fig. 10.14.
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch10 page 478
In0.16 Ga0.84 As QWs. In both cases, introducing the QWs has increased i sc and
reduced Voc . The latter effect results from the increased dark current, which is
evident from Eq. (10.25).
In the case of the Al0.3 Ga0.7 As host cell, where the host bandgap exceeds
the optimum for solar energy conversion, the net effect of QWs is to increase
the cell efficiency. This is as expected, since QWs added to a wide-gap host cell
reduce its effective bandgap towards the optimum. In the case of the GaAs host,
the efficiency decreases, which is again the result expected simply from arguments
about the optimum bandgap for photoconversion: the addition of a lower-bandgap
material to GaAs will reduce the effective bandgap for absorption, and from detailed
balance arguments this is expected to reduce the efficiency of the solar cell.
However, the results for the GaAs and InP host materials are complicated by
material quality issues. In GaAs the number of QWs that may be added before
strain degrades the device quality is too few to produce an adequate photocurrent
enhancement. In InP, problems of high background doping have made it impossi-
ble to prepare good quality devices for comparison. However, In x Ga1−x As QWs
have been observed to increase the efficiency of a less than perfect InP p-i -n cell
(Zachariou et al., 1996). Strain-balanced GaAsP/InGaAs represents the cleanest
material system in which QWSCs have been grown showing a linear increase in
photocurrent with approximately constant voltage (Bushnell, 2005). As discussed
further in Section 10.5, this material becomes dominated by radiative recombina-
tion at high solar concentration and for a short period set a world-record power-
conversion efficiency, discussed further in Section 10.7.2. However, the record was
subsequently reclaimed by a GaAs homojunction (Green, 2011), so it is therefore
not yet possible to decide on the effect, in practice, of QWs on solar cell efficiency.
However, we can learn something from the effect of QWs on the open-circuit
voltage Voc of test devices. Detailed balance arguments (below) imply that Voc
should be controlled by the absorption edge E a . Therefore we would expect a
decrease E a in the absorption edge to cause a decrease in Voc of the same magni-
tude, and we have seen above that the decrease in E a caused by the introduction of
QWs is accompanied by a reduction in Voc . However, measurements (Fig. 10.16)
show that Voc is less sensitive than expected to the effective bandgap E a of the well
material (Barnham et al., 1996). This is reasonable since it is the host material that
controls carrier injection currents, and hence the population of carriers available
for recombination.
Figure 10.16 Open-circuit voltage against effective bandgap for a series of Al x Ga1−x As/
GaAs QWSCs and Al x Ga1−x As p-i-n cells of different Al fraction. Note how, for the QW
devices, Voc is higher than expected either from the measured dependence of Voc on Al
fraction for Alx Ga1−x As devices (dashed line) or from the theoretical dependence of Voc
on the absorption edge of the QW, the effective bandgap, (dotted line) expected from detailed
balance arguments.
an optically thick QW stack, since background impurity levels limit the width of
depleted i-region that can be grown. As a result, rear surface reflectors, such as
a distributed Bragg reflector (DBR), have been employed to increase the optical
path length in the material (Bushnell et al., 2003). These have resulted in quantum
efficiency levels in excess of 50% in the QW region.
Apart from increasing photogeneration, the effect of introducing a high-quality
reflector into the solar cell enables re-absorption of radiative recombination to take
place, an effect known as photon recycling (Marti et al., 1997). In a high-quality
material such as strain-balanced GaAsP/InGaAs, recombination is dominated by
radiative processes at high current levels (equivalent to 200× solar concentration),
and the introduction of reflecting stacks in devices made from such materials was
observed to reduce the electrical dark current of the solar cell (Johnson et al., 2007).
The observed reduction in recombination rate corresponds to a 0.3% increase in
absolute efficiency, yet in principle a gain of 2% in absolute efficiency is possible
through photon recycling.
In addition to trapping light inside the QWSC, it is also possible to manipulate
the emission profile for efficiency gain. In a bulk semiconductor and unstrained
quantum well, radiative recombination is roughly isotropic, but under compres-
sive biaxial strain, the radiative recombination becomes directional, favouring the
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch10 page 480
where jinc is the incident photon flux from the Sun, radiating at temperature Ts , j
is the radiative flux within the cell at temperature Tc with quasi-Fermi level sep-
aration qV (Eq. (10.14)), and F
and Fe are geometrical factors giving the solid
angles for solar photon absorption and emission. For a black-body Sun at 5800 K,
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch10 page 481
the available power i (V )V has a maximum at a bandgap of around 1.3 eV. When
applied to the case of a QWSC of effective bandgap E a this approach gives for the
incremental photocurrent and dark current
Eg
i ph = F
b(E, Ts )a(E) d E
Ea
Eg
i Dk (V ) = Fe j (E, Tc ,
µF )a(E) d E (10.29)
Ea
where a(E) is the probability of photon absorption in the MQW, i.e. its spectral
response. Now for the optimum QWSC a(E) = 1 for all E > E a and, if the quasi-
Fermi level separation µF in the QW is equal to qV, then Eq. (10.26) becomes
identical to Eq. (10.28), and the optimum QWSC will be identical to the optimum
single-bandgap homojunction cell.
There has been some debate about whether the detailed balance theory applies
to the QWSC in practice (Corkish and Honsberg, 1997; Anderson, 1995; Araujo
et al., 1994). Measurements of radiative recombination currents from biased single-
QW test cells suggest that µF is smaller in the QW than in the surrounding host
material. Irreversible carrier escape from the QW under the small electric field
present at the operating point has been suggested as a reason for this (Nelson
et al., 1995; Corkish and Honsberg, 1997). Interestingly, the same effect persists
under illuminated conditions (Ekins-Daukes et al., 2003), lending weight to the
hypothesis of the breakdown of quasi-thermal equilibrium at quantum well length
scales. Microscopic non-equilibrium theories for the QWSC are beginning to reveal
the carrier dynamics present in these systems (Aeberhard, 2010).
Another interesting idea is the possibility of exploiting ‘hot’ carrier effects
(Ross and Nozik, 1982) in QWSCs. At high carrier densities the relaxation of
excited carriers to the band edge can be slowed down by quantum confinement in
a QW. The carrier populations then appear to have a higher effective temperature
than the lattice, and recombination is reduced. Retarded relaxation has already been
observed in QW photoelectrodes (Rosenwaks et al., 1993), with the potential to
reach efficiencies in excess of 50% (Le Bris and Guillemoles, 2010). Some attempts
have been made to design hot-carrier superlattice solar cells (Hanna et al., 1997).
10.7 Applications
Because QWSCs are as costly to produce as high-efficiency III–V homojunc-
tion cells, we may expect them to be interesting only in those applications
where III–Vs are preferred. At the present time that means space, concentrator
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch10 page 482
8
7
6
Efficiency/%
5
4
3
2
1
0
0 10 20 30 40 50 60 70 80 90 100 110
Temperature/C
is not fully understood, clearly the efficiency of carrier escape from the QWs will
increase, or remain at unity, as T is increased. Faster carrier escape is likely to
reduce the probability of recombination in the QWs, and so offset the effect of the
decreasing QW bandgap.
Very high peak performance can be obtained from QWSCs under concen-
trated sunlight. Figure 10.18 shows the increase in efficiency of a GaAsP/InGaAs
quantum well solar cell measured under concentrated sunlight, retaining the tem-
perature stable at 25◦ C. A peak efficiency of 28.3% at 535 Suns was measured
under AM 1.5 D.
10.7.3 Thermophotovoltaics
In thermophotovoltaics, low-bandgap photovoltaic cells are used to produce elec-
tricity from the long-wavelength radiation emitted by a hot (2000–3000◦C) source.
The source is usually provided through fossil fuel combustion in a combined heat
and power system. Often a selective emitter is used to reabsorb the very low energy
photons and reemit them at higher energies to prevent heating. The reshaped spec-
trum is concentrated around certain bands characteristic of the emitter. For such
a spectrum, control of the bandgap of the PV cell is essential for good power
conversion efficiency. The flexibility of bandgap makes QWSCs of great interest
for TPV. It is also possible that Auger recombination, a longstanding problem in
low-bandgap solar cells, is suppressed in the QW device.
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch10 page 484
Figure 10.18 Single-junction 1.2 mm2 GaAsP/InGaAs quantum well solar cell efficiency
measured as a function of concentration. Data courtesy of Quantasol Ltd, measured at the
Fraunhofer Institute for Solar Energy Systems, May 2009.
Quantum well solar cells in InGaAsP/In x Ga1−x As have already been shown
to produce a higher Voc than the comparable Inx Ga1−x As homojunction cell (Grif-
fin et al., 1997; Connolly and Rohr, 2003). For low-temperature emitters, an
In x GaAs/In y GaAs strain-balanced combination grown on InP has demonstrated
absorption thresholds up to 1.95 µm (Rohr et al., 2006).
10.8 Conclusions
We have reviewed the use of novel QW semiconductor heterostructures in solar
cells. Quantum well structures are of interest as a means of enhancing the photocur-
rent and efficiency of crystalline solar cells. Photocurrent enhancement has been
demonstrated in a range of materials and is well understood. Efficiency enhance-
ment has been observed in materials whose bandgap is larger than the optimum for
solar energy conversion. In materials of bandgap close to the optimum, experimen-
tal tests on QW cells of equivalent quality to homojunction cells have not yet been
possible. Nevertheless there is some evidence that the effect of QWs in increasing
recombination within the device is smaller than expected from arguments based on
a quasi-thermal equilibrium distribution of carriers. If this is true under operating
conditions, then higher efficiencies may also be available with optimum-bandgap
cells.
Quantum well structures have the advantages over homojunction cells that the
effective bandgap can be controlled by tuning the width of the QW, rather than
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch10 page 485
by varying the composition of a bulk alloy. This means that QWs may produce
better cells of better material quality than bulk alloys when particular bandgaps are
required. This is relevant for high-efficiency tandem cells and for thermophoto-
voltaic cells, and QW structures are being researched for both these applications.
A further important advantage is that QW structures have a better response to
temperature and consequently are expected to perform better under concentrated
light.
Some of the major challenges that remain are: to find and verify a theoretical
explanation for the observed dark currents and Voc behaviour; to establish whether
the suppressed recombination behaviour observed in the dark occurs under solar
cell operating conditions; and to prepare an optimum-bandgap QWSC of equivalent
quality and superior efficiency to a GaAs homojunction solar cell. More generally,
work on QW structures has stimulated a range of new ideas about the role of
quantum nanostructures in photovoltaics and the limits to efficiency of solar cells.
References
Aeberhard E. (2010), ‘Spectral properties of photogenerated carriers in quantum well solar
cells’, Solar Energy Mater. Solar Cells 94, 1897–1902.
Adams A. (1986), ‘Band-structure engineering for low-threshold high-efficiency semicon-
ductor lasers’, Electron Lett. 22, 249–250.
Adams J. G. J., Browne B. C., Ballard I. M., Connolly J. P., Chan N. L. A., Ioannides A.,
Elder W., Stavrinou P., Barnham K. W. J. and Ekins-Daukes N. J. (2011), ‘Recent
results for single-junction and tandem quantum well solar cells’, Progr. Photovoltaics
19, 865–877.
Afshar M., Sadewasser S., Albert J., Lehmann S., Abou-Ras D., Marron D. F., Rockett
A. A., Räsänen E. and Lux-Steiner M. C. (2011), ‘Chalcopyrite semiconductors for
quantum well solar cells’, Adv. Energy Mater. 1, 1109–1115.
Anderson N. G. (1995), ‘Ideal theory of quantum-well solar cells’, J. Appl. Phys. 78,
1850–1861.
Araujo G. L. and Marti A. (1994), ‘Absolute limiting efficiencies for photovoltaic energy
conversion’, Solar Energy Mater. Solar Cells 33, 213–240.
Araujo G. L., Marti A., Ragay F. W. and Wolter J. H. (1994), ‘Efficiency of multiple quantum
well solar cells’, Proc. 12th. European Photovoltaic Solar Energy Conf., H. S. Stephens
& Associates, Bedford, pp. 1481–1484.
Barnes J. M. (1994), ‘An experimental and theoretical study of GaAs/InGaAs quantum well
solar cells and carrier escape from quantum wells’, PhD Thesis, University of London.
Barnes J., Nelson J., Barnham K. W. J., Roberts J. S., Pate M. A., Grey R., Dosanjh S. S.,
Mazzer M. and Ghiraldo F. (1996), ‘Characterization of GaAs/ InGaAs quantum wells
using photocurrent spectroscopy’, J. Appl. Phys. 79, 7775–7777.
Barnham K., Ballard I., Barnes J., Connolly J., Griffin P., Kluftinger B., Nelson J., Tsui
E. and Zachariou A. (1997), ‘Quantum well solar cells’, Appl. Surf. Sci. 113/114,
722–733.
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch10 page 486
Barnham K., Connolly J., Griffin P., Haarpaintner G., Nelson J., Tsui E., Zachariou A.,
Osborne J., Button C., Hill G., Hopkinson M., Pate M., Roberts J. and Foxon T. (1996),
‘Voltage enhancement in quantum well solar cells’, J. Appl. Phys. 80, 1201–1206.
Barnham K. W. J., Braun B., Nelson J., Paxman M., Button C., Roberts J. S. and Foxon C. T.
(1991), ‘Short-circuit current and energy efficiency enhancement in a low-dimensional
structure photovoltaic device’, Appl. Phys. Lett. 59, 135–137.
Barnham K. W. J. and Duggan G. (1990), ‘A new approach to high-efficiency multi-bandgap
solar cells’, J. Appl. Phys. 67, 3490–3493.
Bastard G. (1988), Wave Mechanics Applied to Semiconductor Heterostructures, Editions
de Physique, Les Ulis.
Bauhuis G. J., Mulder P., Haverkamp E. J., Schermer J. J., Bongers E., Oomen G., Koestler
W. and Strobl G. (2010), ‘Wafer reuse for repeated growth of III–V solar cells’, Progr.
Photovoltaics 18, 155–159.
Bushnell D. B., Ekins-Daukes N. J., Barnham K. W. J., Connolly J. P., Roberts J. S., Hill G.,
Airey R. and Mazzer M. (2003), ‘Short-circuit current enhancement in Bragg stack
multi-quantum-well solar cells for multijunction space cell applications’, Solar Energy
Mater. Solar Cells 75, 299–305.
Bushnell D. B., Tibbits T. N. D., Barnham K. W. J., Connolly J. P., Mazzer M., Ekins-
Daukes N. J., Roberts J. S., Hill G. and Airey R. (2005), ‘Effect of well number on the
performance of quantum-well solar cells’, J. Appl. Phys. 97, 124908–124908-4.
Chaffin R. J., Osbourn G. C., Dawson L. R. and Biefeld R. M. (1984), ‘Strained superlattice
quantum well multijunction photovoltaic cell’, Proc. IEEE Photovoltaic Specialists
Conf., IEEE, New York, NY, pp. 743–746.
Conibeer G., Green M. A., Konig D., Perez-Wurfl I., Huang S., Hao X., Di D., Shi L.,
Shrestha S., Puthen-Veetil B., So Y., Zhang B. and Wan Z. (2010), Progr. Photovoltaics
19, 813–824.
Connolly J. P., Barnham K. W. J., Nelson J., Griffin P., Haarpaintner G., Roberts C., Pate
M. and Roberts J. S. (1995), ‘Optimisation of high efficiency Al x Ga1−x As MQW
solar cells’, Proc. Int. Solar Energy Society 1995 Solar World Congress, Harare, Zim-
babwe, http://www.researchgate.net/publication/235764284_Optimisation_of_high_
efficiency_AlxGa1-xAs_MQW_solar_cells.
Connolly J. and Rohr C. (2003), ‘Quantum well cells for thermophotovoltaics’, Semicond
Sci. Tech. 18, S216–S220.
Connolly J. P. (1998), Private communication.
Corkish R. and Green M. (1993), ‘Recombination of carriers in quantum-well solar-cells’,
Conf. Record 23rd. IEEE Photovoltaic Specialists Conf., IEEE, New York, pp. 675–680.
Corkish R. and Honsberg C. B. (1997), ‘Dark currents in double-heterostructure and
quantum-well solar cells’, Conf. Record 26th. IEEE Photovoltaic Specialists Conf.,
IEEE, New York, pp. 923–926.
Ekins-Daukes N. J., Ballard I., Calder C. D. J., Barnham K. W. J., Hill G. and Roberts J. S.
(2003), ‘Photovoltaic efficiency enhancement through thermal up-conversion’, Appl.
Phys. Lett. 82, 1974–1976.
Ekins-Daukes N. J., Barnham K. W. J., Connolly J. P., Roberts J. S., Clark J. C., Hill G. and
Mazzer M. (1999), ‘Strain-balanced GaAsP/InGaAs quantum well solar cells’, Appl.
Phys. Lett. 75, 4195–4197.
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch10 page 487
Ekins-Daukes N. J., Kawaguchi K. and Zhang J. (2002), ‘Strain-balanced criteria for multi-
ple quantum well structures and its signature in X-ray rocking curves’, Cryst. Growth
Des. 2, 287–292.
Farrell R. M., Neufeld C. J., Cruz S. C., Lang J. R., Iza M., Keller S., Nakamura S., Denbaars
S. P., Mishra E. K. and Speck J. S. (2011), ‘High quantum efficiency InGaN/GaN
multiple quantum well solar cells with spectral response extending out to 520 nm’,
Appl. Phys. Lett.98, 201107–201110.
Freundlich A., Fotkatzikis A., Bhusal L., Williams L., Alemu A., Zhu W., Coaquira J. A.
H., Feltrin A. and Radhakrishnan G. (2007), ‘Chemical beam epitaxy of GaAsN/GaAs
multiquantum well solar cell’, J. Vac. Sci. Technol. B 25, 987–990.
Freundlich A., Rossignol V., Vilela M. F. and Renaud P. (1994), ‘InP-based quantum well
solar cells grown by chemical beam epitaxy’, Conf. Record IEEE First World Conf. on
Photovoltaic Energy Conversion, IEEE, New York, pp. 1886–1889.
Freundlich A. and Serdiukova I. (1998), ‘Multi-quantum well tandem solar cells with effi-
ciencies exceeding 30% AM0’, Proc. 2nd World Conf. Photovoltaic Energy Conver-
sion, Vienna, p. 3707.
Goradia C., Clark R. and Brinker D. (1985), ‘A proposed GaAs based superlattice solar cell
structure with high efficiency and high radiation tolerance’, Proc. IEEE Photovoltaic
Specialists Conf., IEEE, New York, NY, pp. 776–781.
Green M. A. (2003), Third Generation Photovoltaics, Springer, New York.
Green M. A., Emery K., Hishikawa Y. and Warta W. (2011), ‘Solar cell efficiency tables
version 37’, Progr. Photovoltaics 19, 84–92.
Griffin P., Ballard I., Barnham K., Nelson J. and Zachariou A. (1997), ‘Advantages of
quantum well solar cells for TPV’, in Coutts T. J., Allman C. S. and Benner J. P.,
(eds), Thermophotovoltaic Generation of Electricity, American Institute of Physics,
New York.
Griffin P., Barnes J., Barnham K. W. J., Haarpaintner G., Mazzer M., Zanotti-Fregonara C.,
Grunbaum E., Olson C., Rohr C., David J. P. R., Roberts J. S., Grey R. and Pate M.
A. (1996), ‘Effect of strain relaxation on forward bias dark currents in GaAs/InGaAs
multiquantum well p-i-n diodes’, J. Appl. Phys. 80, 5815–5820.
Hanna M. C., Lu Z. H. and Nozik A. J. (1997), ‘Hot carrier solar cells’, in Future Generation
Photovoltaic Technologies — First NREL Conf., McConnell R. D. (ed.), American
Institute of Physics, New York, pp. 309–316.
Hirst L. C. and Ekins-Daukes N. J. (2011), ‘Fundamental losses in solar cells’, Progr.
Photovoltaics 19, 286–293.
Hovel H. J. (1975), Semiconductor and Semimetals, Volume 11 — Solar Cells, Academic
Press, London.
Johnson D. C., Ballard I. M., Barnham K. W. J., Connolly J. P., Mazzer M., Bessiere A.,
Calder C., Hill G. and Roberts J. S. (2007), ‘Observation of photon recycling in strain-
balanced quantum well solar cells’, Appl. Phys. Lett. 90, 213505–213508.
Le Bris A. and Guillemoles J.-F. (2010), ‘Hot carrier solar cells: achievable efficiency
accounting for heat losses in the absorber and through contacts’, Appl. Phys. Lett. 97,
113506–113506-3.
Lee K.-H., Barnham K. W. J., Connolly J. P., Browne B. C., Airey R. J., Roberts J. S.,
Führer M., Tibbits T. N. D. and Ekins-Daukes N. J. (2012), ‘Demonstration of photon
coupling in dual multiple-quantum-well solar cells’, IEEE J. Photovoltaics 2, 68–74.
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch10 page 488
CHAPTER 11
CONCENTRATOR SYSTEMS
IGNACIO LUQUE-HEREDIA
Compañía Española de Alta Eficiencia Fotovoltaica BSQ Solar, SL
Madrid 28040, Spain
[email protected]
and
ANTONIO LUQUE
Instituto de Energía Solar, Universidad Politécnica de Madrid
Madrid 28040, Spain
[email protected]
Porque el que no sabe mas que las palabras sin saber el fundamento que la regla tiene:
siguen se le muchos daños/y hallandose en ellos/no sabe ni alcança de donde le vienen.
(Because much harm will come to him who only knows the words; without knowing the
foundation of the rule, he does not know and cannot reach where they come from.)
Pedro de Medina, Regimiento de Navegación, fo. 26, 1563.
11.1 Introduction
Historically, solar cells have been considered to be expensive. A potential way of
reducing their cost is casting onto them a higher light intensity than is available
naturally. For this solar concentrators are used. Concentrators are optical elements
that collect the Sun’s energy in a certain area and redirect it onto the solar cells.
Obviously the collecting optical element has to be cheaper per unit area than the
solar cell, a necessary although not sufficient condition to render the concentrated
light system less expensive than an unconcentrated one.
The rapid reduction of the cost of silicon solar cells and the irruption of thin-
film cells in the market has weakened the convincing strengths of this approach.
However, another factor that has revitalised the interest in concentrated photo-
voltaics (CPV) has appeared, namely the fabulous potential of multijunction (MJ)
cells for high efficiency under concentration. This fact is perspicuous in the very
widely used National Renewable Energy Laboratory (NREL) chart in Fig. 11.1.
In this we see that the efficiency of triple-junction cells under concentration is not
491
September 17, 2014 11:4
492
9in x 6in
b1809-ch11 page 492
Figure 11.1 Champion laboratory cell efficiencies for different PV technologies. Courtesy of National Renewable Energy Laboratory, Golden,
CO (www.nrel.gov/ncpv/images/ efficiency_chart.jpg).
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch11 page 493
matched by any other photovoltaic technology. In the last decade their efficiency
has increased at the rate of about 1% (absolute) per year.
The increase of efficiency by concentration has solid theoretical grounds that
are explained later in this chapter. One of the authors has expressed his opinion
that 50% cell efficiency will eventually be achievable (Luque, 2011). This promis-
ing potential overcomes the fact that CPV only uses the direct sunbeam and not
the diffuse component of the radiation, which is harnessed by conventional photo-
voltaics (PV). At the same time the earlier motivation of concentration — reducing
the number and hence the cost of the cells required to produce a given power out-
put — becomes convincing in this context. Only under concentrated sunlight can
the sophisticated and expensive multijunction cells be exploited. But we think that
these high efficiencies have the potential of becoming cheaper than any other PV
technology and, furthermore, than any other energy technology.
The level of irradiance (luminous power flux) at which concentrating cells
operate is very variable. Recalling that for the purpose of solar cell rating, the
standard solar irradiance at the Earth’s surface is 1 kW m−2 , the level of irradiance
in static concentrator cells is in the range 1.5–5 kW m−2, while silicon tracking
concentrators range today from 10 to 500 kW m−2, and for multijunction cells
irradiances between 100 and 1500 kW m−2 are used or envisaged. It is very common
to refer to the irradiance level in ‘Suns’, meaning the number of times the actual
irradiance is higher than the standard solar irradiance. Thus a cell operating at
1500 kW m−2 is said to operate at 1500 Suns.
In this chapter, we shall look briefly first at the early history of CPV, then at
the basic operation of solar cells under concentration to explain the grounds of
the increase of the efficiency with concentration and also the limits of this: why
concentration cannot be increased indefinitely and lead to more efficient operation.
Then we outline the modern multijunction structures and their behaviour.
As regards concentrators, we start with their description. Then we examine
methods for their optical design. The theoretical grounds of concentrator optics are
also presented and the limits on increasing the concentration factor are described.
Following that, we analyse how concentrator cells are mounted and cooled, includ-
ing a quantitative analysis of the cooling.
Concentrating systems usually (but not invariably) need tracking mechanisms
to keep the sunlight focussed on the cells. These constitute an important part of
the cost of a CPV system, so we look at them in some detail in this chapter. Static
concentrators also exist, although they permit only very moderate concentration.
They also collect diffuse sunlight, to different extents. Such systems have not so far
been commercialised and will not be treated in this chapter. For more information,
the reader is directed to Luque (1986 and 1989).
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch11 page 494
Figure 11.2 Photovoltaic concentrator panel fabricated in the late 1970s at Sandia National
Laboratories, Albuquerque, NM.
Next we consider the performance and cost of CPV and the potential for cost
reduction. Finally we discuss what future we foresee for tracking concentrator
technology.
(including the Ramón Areces), cooling of the cells was achieved by an extruded
aluminium multifin heat sink. A closed housing was also provided to protect the
cells from the environment. In some cases, parquets of several lenses were fabri-
cated to facilitate the assembling.
The passive-cooling modified Sandia Labs design was adopted by several
manufacturers in those days, Martin Marietta being the most remarkable. This
organisation installed about 350 kW in Saudi Arabia in the early 1980s (Salim
and Eugenio, 1990). Some other US companies, for instance Alpha Solarco and
Amonix (Garboushian et al., 1996) also continued this concept to develop large
panels in the 15–25 kW range.
While many other designs were considered at that time, the next original
concept that came into reality was the ENTECH concentrator. A plant of 300 kW
and several other smaller ENTECH arrays were installed (O’Neill et al., 1991).
In this concentrator, still commercialised today (mainly for space applications),
the cells are series-connected in a linear row located under an arched Fresnel lens
of linear focus, as shown in Fig. 11.3. The performance of such arched lenses is
strikingly insensitive to their position. The concentration ratio can be up to 20×,
and in some cases screen-printed cells are used. Cooling is again effected with
extruded heat sinks, and a ‘housing’ contains the whole module.
The concentrator array comprises a set of linear cell/lens modules in an ele-
vated east–west oriented frame. Each module can rotate separately to follow the
hour angle, and the frame as a whole can rotate in elevation on two supporting
poles situated at its east and west ends. This is the declination–hour angle tracking
mechanism. Its tracking control is similar to that of the Sandia prototype.
A concept developed jointly by BP Solar and IES-UPM, in an EU joint project
in which the University of Reading and ZSW-Stuttgart also participated, was the
EUCLIDES concentrator (Sala et al., 1996) shown in Fig. 11.4. The rationale
Figure 11.3 ENTECH concentrator array. This has axial focus curved lenses, tracking the
hour angle in a frame that tracks the Sun’s elevation angle.
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch11 page 496
Figure 11.4 View of the one-axis tracking EUCLIDES concentrator. The cells, encapsu-
lated inside modules, are the darker strip that can be seen in the figure (there is another strip
underneath). There are two separated asymmetric parabolic mirrors illuminating two rows
of cells.
behind this design was to have a high voltage output, able to feed an inverter without
the use of an intermediate transformer. The output needed for this purpose was 600–
650 V DC (direct current), this requiring many cells (∼1400) to be connected in
series, which resulted in very long arrays (84 m in the commercial version). The
arrays should therefore be horizontal and have a one-axis horizontal tracking. This
implies linear mirror optics, because linear lenses change their focal distance when
the Sun is at inclined angles.
Subject to these specifications, the concept of cell housing was no longer valid.
Instead, cells were encapsulated in receiving modules inspired by flat module
encapsulation, which provided excellent environmental protection. The receiver
module was formed of an aluminium tray, on which the cells were stuck on a layer
of material that had to be both an electric insulator and a good thermal conductor.
The cells were electrically connected in series and covered with glass. The inside
of this module was filled with a transparent resin.
The concentrating optics were made of mirrors instead of the Fresnel lenses
used previously in all (successful) photovoltaic concentration designs. The mirrors
were parabolic in outline, and their profiles were optimised using non-imaging
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch11 page 497
optics (Luque, 1980), so as to allow for the highest manufacturing errors for a given
level of concentration (geometrical concentration of 32×). The focal distance,
position of the axis, useful arch within the infinite parabola, and the receiver (cell)
angle, were also optimised.
Since this mirror profile was not available in the market, and because of the
precise profile that was required (which we were not sure could be achieved with
glass), a new fabrication technology was developed using a weather-protected 3M
silvered layer laminated onto a thin aluminium sheet that was subsequently formed
to the desired shape with high precision and then stuck to a supporting aluminium
frame.
Cell cooling was provided in the EUCLIDES module by means of an
aluminium-finned heat sink, the aluminium sheets being held together by a core
stuck onto the aluminium tray of the modules. Cooling by natural convection is
more effective in vertical fins than in the inclined fins used in the two-axis tracking
configuration, so less aluminium is required for the same cooling performance.
The structure, which held two rows of cells and mirrors as shown in Fig. 11.4,
consisted of a long horizontal reticulated beam rotating on a large centrally placed
wheel resting on two smaller wheels. This configuration also provided the tracking
mechanisms. Two additional passive supports provided the vertical reaction to the
weight at the beam ends. The tracking control was provided by a microprocessor
that calculated the correct aiming of the system at any moment of the year, based on
astronomic data. However, in order to account for inaccurate module positioning,
a feedback based on system power output maximisation was occasionally used,
generating a table of corrections for the astronomic calculations.
The EUCLIDES concentrator used crystalline silicon cells of the laser-grooved
buried contact (LGBG) type, a concept that had been developed by the University
of New South Wales. These cells (also called SATURN, after the name of the indus-
trialisation project of BP Solar) had proven high efficiencies. These cells were, at
the time, very convenient for use in concentrator systems because, while their 1 Sun
version was more expensive than those used in conventional c-Si modules, when
used as 125×125 mm2 concentrator cells and designed for 30×, they could be sold
at 10–12 each, thus becoming economically very attractive to the manufacturer.
In their concentrator version, LGBG cells showed a highly homogeneous voltage
along their metallisation and a low surface recombination, which allowed them to
reach an average of 18.5% efficiency at 30× and up to 20% in small (1 cm2 ) cells
at 100× (Bruton et al., 1994).
The first prototype of this concentrator technology was installed in Madrid
at IES-UPM in 1995. It proved to have an efficiency of 10.8% close to noon on a
typical summer day which, extrapolated to the then-proposed concentrator standard
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch11 page 498
Figure 11.5 Seven of the fourteen arrays constituting the EUCLIDES demonstration plant
installed in Tenerife by the ITER, the IES and BP Solar.
test conditions — direct radiation of 800 W m−2 and cell temperature of 25◦ C —
resulted in an efficiency of 14.4% (Luque et al., 1997). This was achieved at a
lower cost than a typical flat-plate PV power plant of the same rating at that time.
Following the Madrid prototype, a 480 kW EUCLIDES demonstration plant
was built in Tenerife (see Fig. 11.5) through the joint initiative of BP Solar, the
Instituto de Tecnología y Energías Renovables (ITER) in Tenerife and IES-UPM.
This became the biggest CPV plant of its time. However, several problems that
appeared in the series production of the receiver, some overestimation of the con-
centrator benefits, and finally the merger of BP with Amoco (which resulted in the
consolidation of their respective solar subsidiaries BP Solar and Solarex), led to
the abandonment of concentrator projects by these companies.
Other smaller CPV prototypes were developed using combinations of the
above designs. For instance, a point-focus concentrator in a declination–hour angle
frame configuration like that of ENTECH was developed by Midway Laboratories.
Other less fully developed ideas have also been published. For instance, among
the most original suggested was a large reticulated platform resting on wheels or
even floating on water to provide the azimuth tracking. On this platform, rows of
arrays would be installed, each with its own elevation tracking mechanism (Alarcón
et al., 1982).
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch11 page 499
These developments and others, which appeared in the last two decades of
the past century, constituted the ‘first wave’ of photovoltaic concentrators, set-
ting the foundations of CPV and paving the way for current CPV start-ups and
developments.
After this overview of the historical background of CPV, which has also intro-
duced the different elements of a concentrator system by describing several real
examples, we proceed in the next sections to further analyse them one by one and
in greater detail.
Excluding the term in R — the series resistance — ηfill slowly increases with Voc
and therefore with the flux intensity. However, at high currents, when the ohmic
drop increases, the fill factor decreases and this becomes the main factor causing
an efficiency decrease. Obviously this implies that the series resistance must be
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch11 page 500
very small in concentrator cells. This is perhaps the most important difference as
compared with the non-concentrator solar cells.
As a consequence of all this, the solar cell efficiency ηmp
Iph
ηmp = Voc ηfill = SL Voc ηfill (11.2)
PL
increases with concentration until the moment when the ohmic drop causes a reduc-
tion, as seen in Fig. 11.6 (single-junction cell labelled 1J IES-UPM). Here SL is the
cell photosensitivity (approximately constant) and PL is the luminous flux. A very
easy rule is that the highest efficiency is achieved when Iph RS ≈ Vth (Luque, 1989).
This behaviour, although ruled by a more complex expression, is also present (for
the same fundamental reasons) in MJ solar cells, as can also be seen in Fig. 11.6.
A very important feature of the concentrator solar cells is their thermal
behaviour. At higher temperatures the bandgaps decrease and therefore the cur-
rent tends to increase, but this is a very small effect — the relative increase of the
short-circuit current in silicon is only 3 × 10−4 K−1 . The most important thermal
effect in a solar cell is the voltage reduction, which is governed by the expression
dV E g /q − V − I RS
=− (11.3)
dT T
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch11 page 501
further assumed that all non-radiative recombination processes are suppressed and
that radiative recombination is only produced by radiation that escapes through the
cell surfaces. Thus they established a balance equation of the photons entering and
escaping the semiconductor and the electron-hole pairs produced and extracted.
No mention is made of the properties of the semiconductors — excepting the
bandgap — or of the need of any p-n junction.
The calculations were repeated by Araújo and Martí (1994) for the case of solar
radiation impinging on the solar cell isotropically rather than within the natural
cone of direct solar radiation. As we will see in the next section, this condition
corresponds to the maximum possible sunlight concentration at the Earth’s surface.
In this case the maximum possible efficiency is 40.7% for black-body radiation
at 6000/300 K (Sun/ambient) temperatures, as shown in Fig. 11.7, together with
cases for other spectra (direct and global) and concentration levels.
However, if the design principles of the solar cells are not considered, and
only thermodynamic considerations are taken into account, the limiting efficiency
of a solar energy converter was calculated by Landsberg and Tonge (1980) and is
Figure 11.7 Shockley–Queisser efficiency limit for an ideal solar cell versus bandgap
energy for: a) unconcentrated 6000 K black-body radiation (1595.9 Wm−2 ); b) fully con-
centrated 6000 K black-body radiation (7349.0 × 104 W m−2 ); c) unconcentrated AM 1.5-
Direct13 (767.2 W m−2 ); d) AM 1.5 Global13 (962.5 W m−2 ). Reprinted with permission
from Araújo and Martí (1994), © Elsevier.
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch11 page 504
given by
4
4 Tamb 1 Tamb
ηLandsberg = 1 − + (11.4)
3 TSun 3 TSun
Assuming the Sun to be a reservoir of photons at 6000 K (TSun ) and the Earth a cold
reservoir at 300 K (Tamb ), this results in a maximum 93.33% maximum conversion
efficiency (Section 2.4 of this book gives a detailed treatment).
The main reason a solar cell, even if ideal as in the SQ assumptions, is so
far from the Landsberg efficiency is that it does not use the whole solar spectrum
but only those photons of energy higher than the bandgap. A second reason is that
entropy is generated by the cooling of the electron-hole pairs produced by photons
with energy higher than the bandgap.
The straightforward way to overcome the fundamental limitation of the SQ
single-bandgap cell is to use several solar cells of different bandgaps to convert
photons of different energies. A simple configuration to achieve this is vertical
stacking of the cells so the uppermost cell has the highest bandgap, and lets the
photons with energy less than its bandgap pass through to the cells below. The last
cell in the stack is the one with the narrowest bandgap. In this way the entire solar
spectrum can be used and the upper cells can act as a filter for the lower ones,
each being illuminated by a narrower range of photon energies, thus reducing
the production of entropy. For an infinite number of ideal cells, as per the SQ
definition, the maximum efficiency of this arrangement is 86.8% under maximum
concentration, for a 6000/300 K Sun/ambient temperature.
The most common practical realisation of this concept is the so-called mono-
lithic MJ cell, consisting in a stack of solar cells of different semiconductor mate-
rials, along with tunnel diodes to interconnect them in series, all grown on a single
substrate. III–V compound semiconductors of a variety of materials can be grown
epitaxially with very high quality, if all the materials have a similar lattice con-
stant (although the so-called metamorphic cells allow for some lattice mismatch to
achieve a better bandgap combination). Figure 11.8 shows a schematic of a stack
of three cells built on a Ge substrate; the figure caption shows the roles of the dif-
ferent layers. The so-called window layers are intended to prevent recombination
at the surface or in the tunnel junctions. The fraction of Group III elements in each
layer is adjusted to obtain the proper lattice constant and, if possible, the desired
bandgap. As noted in the figure, the data are illustrative and each manufacturer has
their own recipe.
As noted above, the different cells in a monolithic MJ device are series-
connected, and this introduces an additional constraint, because the photogenerated
current of the stack will be the smallest of the currents generated by each cell.
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch11 page 505
The best situation is when all the cells produce the same photogenerated current.
This restricts the efficiency of a series-connected stack with respect to that of an
unconstrained stack. For an infinite number of cells, the efficiency limit is the same
(Tobías and Luque, 2002). However, in the case of three cells (the configurationthat
is commercially available today), the efficiency limit for the unconstrained stack
is 63.6% compared with 63.0% for the series-connected stack, where conditions
of maximum concentration (about 46000 Suns) and Sun/ambient temperatures of
5762 K/298.15 K are assumed in both cases. The Sun temperature considered here
is that deriving from the extra-terrestrial spectrum and the ambient temperature
is 25◦ C. These are some of the commonly used standard measuring conditions.
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch11 page 506
Optimal bandgaps for the series-connected case are 1.75, 1.10, 0.58 eV, but opti-
mal bandgaps at lower concentration are higher, and efficiency is lower.
Two main strategies are followed for the growth of the epitaxial cells: the
lattice-matched (LM) approach and the metamorphic (MM) approach. In the first
case, the lattices of the different cells are as closely matched as possible. That is
why a small proportion of In (1%) is added to the GaAs, allowing close matching
with the Ge substrate. The drawback is that this procedure does not admit optimal
bandgaps for ternaries of the elements under consideration (Ga, In, P and As). This
is evident in Fig. 11.9 (Guter et al., 2009), where, for the lattice-matched (LM)
cell, the bandgaps of the two upper cells are a bit too large. In this scheme, the
upper III–V cells give too little current as compared with the Ge cell.
In the metamorphic approach the bandgaps are better adjusted, but at the
expense of permitting the formation of threading dislocations, as shown in
Figure 11.9 Detailed balance calculations for the efficiency of different triple-junction
solar cell structures under the AM 1.5D ASTM G173–03 spectrum at 500 kW m−2 and
298 K. The grey haze, whose contours are projected in the three Cartesian planes, rep-
resents bandgap combinations, which allow efficiencies from 60.5% to 59.0% and hence
mark the optimum. Five specific triple-junction solar cell structures are shown: the lattice-
matched Ga0.5 In0.5 P/Ga0.99 In0.01 As/Ge (LM), two metamorphic GaInP/GaInAs/Ge (1.8,
1.29, 0.66 eV for MM1) and (1.67, 1.18, 0.66 eV for MM2), as well as two inverted meta-
morphic GaInP/GaInAs/GaInAs (1.83, 1.40, 1.00 eV for Inv1) and (1.83, 1.34, 0.89 eV for
Inv2) devices. Reprinted with permission from Guter et al. (2009), © AIP.
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch11 page 507
Figure 11.10 a) Lattice-matched and metamorphic three-junction cells grown on Ge. Dis-
locations are shown in the stressed layers; b) inverted metamorphic cell grown on GaAs,
detached and bonded to carrier; c) bifacial epitaxy metamorphic cell. Reprinted with per-
mission from Wojtczuk et al. (2010), © IEEE.
Fig. 11.10. This reduces the cell quality and may balance out the advantages of
better bandgap matching.
Very good results have been obtained with the present structures, as detailed by
Yamaguchi in Chapter 7. Several three-junction cells have reached efficiencies of
well over 40%; in late 2012, the world record was held by Solar Junction’s LM three-
junction InGaP/GaAs/GaInNAs cell, which is 44% efficient at 947 Suns. As for
four-junction cells, in autumn 2013 the record was held by Fraunhofer/Helmholtz
Centre’s cell, reported as 44.7% efficient under 297 Suns. These efficiency mea-
surements have been made at certified laboratories. However, it is worth noting
that uncertainty in the measurements is given as ±2.5 absolute percentage points
(Green et al., 2010).
Further increase in the efficiency of MJ cells may come through the use of
quaternary compounds that provide an additional degree of freedom to tune lattice
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch11 page 508
of the field are big compared with the wavelength. Then the concept of the light ray
is defined as the curve that is always normal to the wavefronts. The light ray (light
path) between any two points will have to satisfy Fermat’s principle, this being a
variational formulation stating that the length of the optical path followed will be
stationary (dL = 0), i.e. a maximal, a minimal or a saddle point, where the optical
path length between two points is defined as
A
L= n(x, y, z)dl (11.5)
B
where n(x, y, z) is the refractive index of the medium at point (x, y, z) and dl is
the differential length along the light’s path between points A and B. All the laws
of geometric optics can be derived from Fermat’s principle, including Snell’s laws
of reflection and refraction. In fact, minimal path length is the case when a light
ray passes from one medium into another (refraction), or is reflected by a planar
mirror (this was already known by the Greeks, who formulated the reflection law
exactly through Fermat’s principle). However, it can also be a maximum, as occurs
in gravitational lensing.
An alternative formulation of geometrical optics results from the Hamiltonian
equations. A ray, specified as passing through a point (x, y, z) with a direction
given by the unitary vector v̄ of its direction cosines, is represented as a six-
vector (x, y, z, p, q, r ) where ( p, q, r ) = n(x, y, z)v̄ is the vector of the so-called
optical direction cosines of the ray. The Hamiltonian formulation states that the
rays follow trajectories that are solutions of the following system of first-order
differential equations
dx ∂H dp ∂H
= =−
ds ∂p ds ∂x
dy ∂H dq ∂H
= =−
ds ∂q ds ∂y
dz ∂H dr ∂H
= =− (11.6)
ds ∂r ds ∂z
which measures how many rays a manifold has. The theorem of conservation of the
etendue is fundamental to non-imaging optics design and it states that the etendue
is an invariant of any ray manifold as it propagates through an optical system.
In other words, the etendue is independent of the reference surface used for its
calculation.
In concentrator optics design, the bundle of rays that come from a certain light
source, the Sun in the case of solar energy, and impinge on the entry surface of the
optical concentrator is called the input manifold and is denoted by Mi . The manifold
of rays that traverse from the exit surface of the concentrator to the photovoltaic
cell, performing as receiver is denoted by Mo . The collected manifold Mc of rays
is defined as those rays that are connected by the concentrator and belong both to
the input and exit manifold. Also Mo is a subset of the manifold M R composed by
all the rays that can reach the receiver.
Design problems in non-imaging optics can be of several different types. If
the design is intended to couple the input and output manifolds perfectly, the
concentrator obtained will be such that (Mi ) = (Mo ) = (Mc ), and it is
then said to be ideal. If (Mc ) = (Mo ) = (M R ) then the concentrator is
said to be maximal provided it is able to illuminate the receiver isotropically.
A concentrator design with the properties of being both ideal and maximal is known
as optimum. On other occasions the target may be to just have Mi included in Mo ,
(Mc ) = (Mi ) ⊂ (Mo ), in which case we can impose an additional design
condition, which could be that Mc produces a prescribed irradiance distribution
on the receiver, e.g. homogeneous irradiance in photovoltaic cells improves their
conversion efficiency.
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch11 page 511
There are some ray manifolds that are especially useful. For example, the so-
called finite source is that of the rays that connect two parallel concentric planar
discs, both within the same homogeneous optical medium. When one of the discs
is moved away from the other towards infinity, then we talk of an infinite source
and the manifold consists of all the rays that impinge on the other disc with an
incidence angle (with respect to the entry surface normal) less or equal to a certain
constant angle α, named the source’s acceptance angle. It can be proved that the
etendue of an infinite source E inf is
where A is the disc’s surface and n is the refractive index of the surrounding
medium.
The geometrical concentration is defined as the ratio of the area of the entry
surface of the concentrating system to that of the receiver. As we have already
been doing, in a concentrator design the geometric concentration is usually written
followed by an ×, e.g. a 1000× design is a concentrator with an entry surface 1,000
times bigger than the surface of the cell.
In order to better quantify geometrical concentration, let Mi be the ray manifold
of an infinite source with acceptance angle α, impinging on the entry surface of a
concentrator system surrounded by a medium of refractive index n. Also let Mo
be the output manifold, characterised by a homogeneous angular spread β, and
n o the refractive index of the medium surrounding the receiver. Then if the entry
and receiver surface areas are AE and AR respectively, due to the conservation
of etendue the input and output manifolds, E Mi and E Mo , must be equal if the
concentrator perfectly couples these two manifolds (i.e. it is an ideal concentrator):
where the last inequality relates the input and output etendue to the maximum
possible etendue at the receiver when this is illuminated isotropically (β = π/2)
(i.e. it is a maximal concentrator), from which a maximum limit for geometric
concentration Cg can be derived:
AE n2
Cg = ≤ 2 o2 (11.10)
AR n i sin α
This is usually referred to as the thermodynamic limit of concentration, because it
can also be derived by equalling the temperature of an ideal black absorber to that of
the Sun considered a black-body radiator with TS = 5777 K (this is closer to reality
than the 6000 K often used for mnemonics) in accordance with the second law of
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch11 page 512
thermodynamics. In the most common case of the entry surface to the concentrator
being surrounded by air (n i = 1), Eq. (11.10) turns into
n 2o
Cg ≤ (11.11)
sin2 α
This can also be expressed as
where the square of the refractive index of the medium surrounding the receiver
becomes the upper bound to the concentration acceptance product (CAP). This
parameter is frequently used today as a figure of merit of concentrator designs,
showing how close they get to ideality.
If we have a four-parametric ray manifold M, then δ M is defined as its edge-
ray manifold and is the boundary, in topological terms, of that manifold. δ M is
itself a three-dimensional subset of manifold M. One of the most powerful tools of
non-imaging optics design is the so-called edge-ray theorem, stating that if we want
to perfectly couple two manifolds Mi and Mo then it is enough to match their δMi
and δ Mo subsets. This can be expressed as (Mi ) = (Mo ) ⇔ (Mi ) = (Mo ) as
the inverse relationship is also true: if Mi and Mo are coupled by an optical system
then their edge-ray manifolds will also be coupled. This means that the design
process can focus just on edge-ray manifolds, thus reducing by one the dimension
of the problem of coupling four-dimensional manifolds. Even though this theorem
was assumed to be true and has been in use since the beginning of non-imaging
optics, it was not to be proved till the mid-1980s by IES-UPM’s Miñano (1985a,
b and c).
The problem of coupling two four-dimensional manifolds is usually too com-
plex, and in order to make it more attainable the usual simplification consists in
imposing an axis or plane of symmetry, on both the ray manifolds and the optical
system developed, and then working out the design in one of the planar sections.
Then it is a couple of two-dimensional manifolds, m i and m o subsets of Mi and
Mo , that are to be matched. For example, in the case of assuming a rotationally
symmetric optical system the design is carried out in one of its meridian planes —
those that contain the axis of symmetry — where ideal solutions perfectly matching
m i and m o can be obtained. Afterwards the three-dimensional optical system will
be obtained by rotating the two-dimensional resulting design around the symmetry
axis. However, there will be no guarantee that non-meridian rays in the Mi and
Mo manifolds will also be ideally coupled, and the assessment of the behaviour
of these other rays usually is completed using ray-tracing software applications.
Ray-tracing CAD tools, essentially able to compute Snell’s laws for very large sets
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch11 page 513
of light rays as they proceed through different optical media, are very much used
in non-imaging optics to simulate and characterise the performance of resulting
designs.
An important characterisation parameter of CPV optical designs is the total
transmission ratio, defined as
P(Mc )
η= (11.13)
P(Mi )
where P(M) is the power transmitted by the M ray-manifold. Then the optical
concentration Co of the system is defined as
Co = C g η (11.14)
When designing for infinite sources with a certain acceptance angle, trans-
mission can be plotted as a function of the rays’ input angle θ so that we obtain the
transmission angle curve
d P[Mc (θ, θ + dθ )]
T (θ ) = (11.15)
d P[Mi (θ, θ + dθ )]
where d P[M(θ, θ +dθ )] is the infinitesimal power transmitted by rays in manifold
M in the angular differential interval (θ, θ + dθ ). The maximum of T (θ ) is termed
as the optical efficiency, ηo , usually occurring at normal incidence, i.e. θ = 0.
Several different design methods have been developed in the field of non-
imaging optics. The earliest was proposed by Welford and Winston (1978) and
used by them to develop all the family of the so-called CPCs (compound parabolic
concentrators) based on the application of the edge-ray theorem in two dimensions
(Fig. 11.11). The primary prototype of this family of concentrators consists of two
symmetric parabolic mirrors, each focusing edge rays entering the concentrator
with the maximum acceptance angle at the other’s lower rim, which coincides
with the receiver’s edge. In its two-dimensional version, this design performs as
an optimum concentrator. Several other versions of the CPC were later developed,
such as the dielectric-filled CPC (Ning et al., 1987) using lossless total internal
reflection instead of mirrors (which incur losses of ∼10%), or CPC designs for
finite sources or for non-maximal concentrations.
Welford and Winston attempted the extension of their initial edge-ray design
method to three-dimensional systems with their flow-line method, the most remark-
able product of which was the hyperboloid concentrator, also called the trumpet
concentrator (Winston and Welford, 1979a, b). Later, the Poisson bracket method
was developed by Miñano (1985a, b). While this method was specifically con-
ceived for the design of three-dimensional concentrators, it usually requires the
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch11 page 514
Figure 11.11 Winston’s CPC consisting of two symmetric parabolas with an inner mirrored
surface, and with their focal points at each end of the receiver. While the two-dimensional
version performs as an optimum concentrator, the plot on the right shows the transmission
performance of a three-dimensional rotationally symmetric design, with a design acceptance
angle of 16◦ where transmission is brought down to zero in approximately 1◦ .
use of variable refractive index media, this being impractical in most cases. The
main interest of the Poisson bracket method is that it is able to produce ideal three-
dimensional concentrators, thus ultimately proving their theoretical existence. It
also produced a family of ideal two-dimensional concentrators, the compound tri-
angular concentrator (CTC), with two different regions of constant refractive index
(Miñano, 1985c) (see Fig. 11.12).
The Simultaneous Multiple Surface (SMS) method was developed in the early
1990s at IES-UPM (Miñano and González, 1992) as a two-dimensional method,
and has to date produced a long list of different designs. Quite curiously, it was born
while trying to solve a conjecture. It was known that a single refractive surface can
exactly focus a plane wavefront into a point; these sort of surfaces are the so-called
Cartesian ovals. The conjecture, based once again on the edge-ray theorem, was
that two plane wavefronts could be exactly focused into two points, the receiver
edges, by means of consecutive refractions or reflections in two optical surfaces.
In fact the conjecture can be extended into the imaging optics world, by rephrasing
it as a number n of spherical wavefronts generated at object points in the near field
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch11 page 515
(or again plane wavefronts for points at infinity) that are to be focused into n image
points, again by sequential refractions or reflections in n optical interfaces.
No formal proof of the two-surface conjecture has yet been produced, but the
SMS method was developed as an iterative method by which these two sought
optical surfaces could be generated. Several designs have been produced by this
method, generally grouped by families named RR, XR, RX, XX, RXI, where the R
stands for refraction, X for reflection, and I for total internal refraction, which are
listed in the same order as incoming light rays will encounter their two constituent
optical interfaces (Fig. 11.13).
Figure 11.13 a) The RXI (Refraction + Reflection & Total Internal Reflection) concentrator
designed using the Simultaneous Multiple Surface (SMS) method. These light rays refract
at the front surface, then reflect in the mirrored back surface and are finally reflected towards
the cell in the inner side of the front surface. This reflection is mostly total internal reflection,
except for the central part in which a reflective spot is placed to redirect rays arriving at high
angles; b) represents an RX SMS design. In a) and b), the entering edge rays arrive at both
ends of the receiver and at the receiver; thus obeying the edge-ray theorem.
Figure 11.14 Schematic of a Fresnel lens showing its facets. The slopes of a plano-convex
lens (above) are projected into a thinner plate.
For instance, dome- shaped Fresnel lenses as shown in Fig. 11.15 exhibit much
better characteristics.
What we usually desire of the optical design is to achieve the highest possible
concentration with the highest optical efficiency. In addition we wish this efficiency
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch11 page 518
where n o is the index of refraction of the medium surrounding the cell and φS is
the Sun’s angular semi-diameter (about 0.265◦ ).
Even theoretically such concentrations cannot be achieved with the lenses or
mirrors we have described above, because they are not able to provide the necessary
isotropic illumination of the receiver. This isotropy can be increased through the
use of two-stage optics, with a primary optics element (POE) that directly receives
sunlight at its entry surface, and a secondary optics element (SOE), that receives the
concentrated light from the primary, and further conducts it into the PV cell sited
at its exit surface. Apart from increasing concentration in the receiver by enabling
light isotropy gain, secondaries are usually designed to increase homogeneity of
the light flux in the PV cell or to increase the acceptance angle of the overall
concentrating system. In the design terms discussed previously, the theoretical
requirement of a secondary is that it is able to cast ray manifolds from a finite
source, the output of the primary, into the receiver.
While not practically feasible as a POE — where small design acceptance
angles are involved — non-imaging CPCs or their derivatives, such as the dielectric-
filled CPCs or trumpet concentrators, can be used as efficient secondary optics as
acceptance angles for this stage are much bigger. Other simpler devices, such as
the truncated cone (for round cells) or truncated pyramid (for square cells), have
also been used. In these devices the inner walls are reflective — either made from
aluminium or covered with some reflective film — and redirect into the cell those
rays that enter and would fall beyond the cell’s perimeter, and increasing concen-
tration and acceptance angle. However, solid dielectric secondaries can achieve
higher concentrations for the same acceptance angles thanks to the n 2o factor in
Eq. (11.16) being higher than unity. Another successful secondary design along
these lines is the dielectric truncated pyramid, where lossless total internal reflec-
tion is used to provide two to four times additional concentration and at the same
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch11 page 519
time, through the so-called kaleidoscopic effect, produces high homogeneity in the
light flux in the cell.
One of the ways of providing some tolerance in manufacturing, assembly and
Sun-tracking is to design for the collection of rays from a region of the sky larger
than the solar disk, i.e. to increase the optical design’s acceptance angle. The max-
imum concentration achievable by this means can be calculated from Eq. (11.16)
using an angle greater than the Sun’s semi-diameter for φs . Table 11.1 (Luque,
1989) shows the limit of concentration for several families of concentrators for the
case of collecting only the Sun’s semi-diameter, 0.265◦ , and for the more prac-
tical case of an allowance of 1◦ . The concentrator’s primary entry surface (also
called its aperture) is specified through the f-number, which is defined as the ratio
of aperture diameter to the focal distance. This table can be used to estimate the
angular allowance required by a certain design. For instance, the above-mentioned
ENTECH concentrator with arched Fresnel lenses had a geometrical concentration
of 20× (see Table 11.1) and the value in Table 11.1 for an angular tolerance of
1◦ is 26×. Simple proportion gives us an approximate angular allowance for this
concentrator of 1.3◦ . In the table we can observe that linear concentrators achieve
much lower concentrations than point-focus ones. This is because there is no con-
centrating effect in one of the dimensions. For two-dimensional concentrators, the
maximum thermodynamic concentration is only the square root of what appears
in Eq. (11.16). Note that the highest concentration for the Sun’s angular spread is
54× (limited by chromatic aberration) for arched Fresnel lenses, 108× for linear
parabolic mirrors and 376× for point-focus Fresnel lenses. In some cases con-
centrations higher than those shown in Table 11.1 with high efficiency have been
announced; and this is surprising to us as the results in this table are the result of
fundamental limitations.
Table 11.1 also shows how the highest concentrations are achieved through
the integration of secondaries. However, it must be noted that full isotropy in the
receiver is not practical, if incidence angles beyond 65◦ Fresnel reflection on the
cells is regarded as too high. So then again applying the conservation of etendue
principle, the maximum practical gain of a secondary is:
n 2o sin 65◦
C g sec ≤ (11.17)
sin2 αsec
where αsec is the incidence angle at the entry surface of the secondary, which can
be expressed as a function of the f-number as tan αsec = 1/2f. Thus we obtain
Table 11.1 Maximum concentration for concentrators accepting all incident rays.
Angular Angular
acceptance 0.265◦ acceptance 1◦
Concentrator type C g (Cgsec ) Aperture Cg (C gsec ) Aperture
Source: Luque (1989). C g is the overall geometrical concentration (primary optics aperture-to-cell area
ratio); Cgsec is the secondary geometrical concentration (aperture of the secondary-to-cell area ratio),
where the refractive index of the secondary is 1.49, which is that of PMMA. Aperture diameter or width
(for the primary) is the (freely selected) focal distance divided by the number in the table (f-number).
so that, for example in the case of the square flat lens with a secondary with
n o = 1.49 in Table 11.1, we find Cg,sec ≤ 37.9, i.e. lower than that achieving max-
imum overall concentration.
Figure 11.16 shows the results of the analysis of the transmission angle curve,
as defined above, of different common secondaries when placed below a Fresnel
lens with f/1 (Victoria et al., 2009). It can be seen that the rotational CPC provides
the highest acceptance angle.
Another important requirement for concentrator optics is homogeneity of illu-
mination. Strongly inhomogeneous illumination produces much higher than aver-
age local concentrations, leading to losses in the fill factor of the cell and reducing
efficiency. One solution is the above-mentioned kaleidoscope SOE, consisting of
a truncated glass pyramid in which the irradiance is expected to be homogenised
by multiple total internal reflection of the entering rays. However, this solution
substantially reduces the acceptance angle. An interesting solution has recently
been presented by Benítez et al. (2010): the Fresnel–Koehler (FK) concentrator,
which uses the principle of designing the SOE to image the POE onto the cell (see
Fig. 11.17). As the POE is homogeneously illuminated by the Sun it produces a
homogeneous light profile. The POE can be accurately imaged onto the cell with
the SOE. This produces a nice squared homogeneous illuminated area that fits with
the square shape of the cell. The FK SOE has the aspect of a dome and presents an
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch11 page 521
Figure 11.16 Angular transmission curves for the different SOE (secondary optics element)
studied. Lens-to-cell geometrical concentration 1000× and f/1. Reproduced with permission
from Victoria et al. (2009), © OSA.
inactive optical region near the cell that is helpful when sticking the concentrator
to the cell (if this zone is active it gives rise to high losses due to the brimming
glue). As matter of fact in the figure, the FK concentrator is formed of four inde-
pendent POE–SOE couples, all four manufactured together, so this design aspect
is irrelevant for the user. Their acceptance angle is reasonably high, in the range
of 1.1◦ for 1000×.
The optics introduces changes in the spectrum impinging on the cell so that
its optimum design will not be that used today in champion cells designed for a
standard spectrum. Furthermore, chromatic aberrations may cause the illumina-
tion profile at the cell to be different for different spectral ranges. An additional
advantage of the FK concentrator is that it is almost free of chromatic aberrations.
Reflective optics is also used, although much less often. Figure 11.18 shows
the solution of the pioneering CPV company SolFocus, based on non-imaging
optics principles (Winston and Gordon, 2005). Its acceptance angle is rather high
(∼1.4◦ for 500 Suns) but the efficiency is relatively low owing to metal reflec-
tions (governed by the metal absorption) and two Fresnel reflections. Note that the
element over the cell is a homogenising kaleidoscope. Mirrors can bend rays more
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch11 page 522
Figure 11.17 Above: View of the Fresnel–Koehler POE and SOE. This is formed of four
POE in a single Fresnel lens, each one imaged onto the cell by the four SOE elements. The
four SOE elements are cast in the same block. Below: Irradiance distribution on the cell for
the FK concentrator with C g = 625×, f/1, no antireflective (AR) coating on the SOE, when
the Sun is on axis and the solar spectrum is restricted to: a) the top sub-cell range (360–690
nm); and b) the middle sub-cell range (690–900 nm). Reproduced with permission from
Benítez et al. (2010), © OSA.
Figure 11.18 Cassegrainian optics in the SolFocus CPV module. Two reflections, first
in a parabolic dish and then in a hyperbolic mirror with same focus, sends light down to
the receiver block where a TIR kaleidoscope homogenises light reaching the cell below.
Courtesy of SolFocus, Inc.
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch11 page 523
Figure 11.19 Side-cut view of a common concentrator module design, based on a parquet
of Fresnel lenses as primaries.
than lenses; therefore this concentrator is about three times more compact than a
refractive system for the same POE entry aperture.
Figure 11.20 Concentrix’s 500× module, following the all-glass design principle (initially
the side panels were also made of glass and one could look inside the module and check if
it was roughly focused). On the right is an inside view of the module where interconnected
cell substrates can be seen receiving the concentrated sunspot.
every manufacturer and details are usually kept secret. Some have opted for making
a vacuum inside the enclosure, which is hermetically sealed, while others go for
an open approach in which venting holes are only made dustproof through metal
cloth coverings, and rain water is allowed inside provided the cell encapsulation
and wiring is made waterproof. In between these two approaches, manufactur-
ers frequently choose venting solutions based on holes covered with breathable
waterproof Gore-Tex membranes.
Finally, it is interesting to describe, following Sala’s analysis (Sala and Antón,
2011), the impact of cell size on several other module design parameters and what
the range of its optimum value is. For the most usual case of Fresnel primaries,
optical efficiency decreases with f-number and usually this value is set above unity.
Assuming a minimum f-number of one, then we can establish an approximate
relation between cell size and focal distance
AE F2
AC = = (11.19)
Cg Cg
where AC is the cell size; AE is the primary lens surface or, in optical design terms,
the entry aperture surface; F the focal distance; and Cg the geometric concentra-
tion. In other words, the smaller the cell the less the module volume, this meaning a
reduction in the cost of the housing materials and in the module’s requirements for
storage and transportation space, and frequently in the cost of the tracker compo-
nents as well (not only because of a possible reduction in module weight, but also
because the centre of gravity of the array of modules stays closer to the tracking
drive, reducing its load rating). Also, smaller cell size will result in lesser series
resistance and consequently I 2 R ohmic losses, and as will be seen below a smaller
size also enables better heat dissipation. Furthermore, the smaller cells attain a
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch11 page 526
Figure 11.21 Side view of one of the early designs of Daido Steel, a 550× module inte-
grating domed Fresnel primaries where one enclosure side panel has been removed to show
the module’s inside.
higher wafer-filling factor (i.e. the ratio of the total surface of cells made in a wafer
to the total wafer surface). But there are also some disadvantages in going to too-
small cells, for example a lesser fraction of active area per cell, or higher material
losses when dicing the cell wafers, and usually higher complexity in the assembly
process. Figure 11.21 shows a side inner view of one of the early module designs
of the Japanese manufacturer Daido Steel, while Fig. 11.22 shows some examples
of current and recent commercial high concentration PV (HCPV) modules.
Sala’s costing exercise assumed a fixed geometric concentration of 1000× and
20% efficient modules with a design based on a flat Fresnel lens as primary, and
a silo-type secondary. It focuses on two production regimes of 10 MW yr−1 and
30 MW yr−1, which are considered likely in the start-up phase of a CPV production
facility. By exploring cell sizes from 1 mm2 to 100 mm2 the curve in Fig. 11.23
is obtained, which shows a module cost minimum for cells of size between 10–
20 mm2 in area, beyond which module cost increases slowly due to decreasing
efficiency deriving from higher thermal losses.
Figure 11.22 Different commercial CPV modules: a) the US’s SolFocus; b) Spain’s Iso-
fotón; c) Taiwan’s Arima Eco. Courtesy of SolFocus, Inc. Isofotón, SA, Arima Eco Energy
Technologies Corp.
Figure 11.23 Estimated cost estimate of 1000× module based on Fresnel POE and silo
SOE, rated at a DNI of 850 W m−2 in a manufacturing scenario of 30 MW yr−1 . The black
curve assumes fixed efficiency while the red curve takes into account efficiency variations
due to cell size, showing the impact of cooling degradation as the cells grow larger. Courtesy
of Prof. Sala, personal communication.
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch11 page 528
The heat is produced at the cell itself, mainly where light absorption occurs.
This tends to be at the cell front. Thus most of the heat goes through the cell volume
to the heat sink or back plate, and then to the interface with air of this metallic part,
where it is transferred to the ambient. Temperature gradients are the driving force
for this heat flow, so some temperature rise over the ambient cannot be avoided.
In many parts of the path, the heat is transferred by conduction so the heat flow
density is proportional to the gradient of the temperature, the thermal conductivity
κ being the constant of proportionality. This law, formally equivalent to Ohm’s law
of electric conduction, justifies the use of an electric equivalence between heat and
current flow. In this equivalence the heat flow plays the role of the electric current
and the temperature that of the electric potential. For a rod of cross-section A and
thickness d the thermal resistance (K W−1 ) is then
Table 11.2 gives the thermal conductivity of some common materials, and
shows the importance of the encapsulation materials. Copper and aluminium are
good conductors but other metals are poorer. However, there are metals (like iron,
κ = 0.5 W cm−1 K−1 ) with a linear expansion coefficient better matched to that
of solar cell semiconductors than Cu, and with a reasonable thermal conductivity
so that they can be used as the stage on which the cell is bonded. Cells are either
soldered to these metal substrates or attached with heat- and electricity-conducting
glues. In order not to shunt the cell, it has to be electrically isolated from the metal
substrate performing as heat sink by an insulator that should also be a good thermal
conductor. In this respect several solutions are applied, usually coming from the
power electronics world, such as the use of IMS (insulated metal substrate) or DCB
(direct bonded copper) substrates that are essentially printed circuit boards (PCBs)
Thermal Thermal
conductivity/ conductivity/
Material W cm−1 K−1 Material W cm−1 K−1
convection and radiation at the back plate (radiation through two surfaces and
convection through the outer one) with κ R in the radiation term, the right addend,
being Wien’s constant and ε being the surface emissivity.
The temperature difference between the coldest point of the back plate and
the receiver bottom is
AE
Ti − To = −B(n)ηop ln(α Rc ) (11.22)
πwκ
where B(n) is the direct irradiance, Rc is the cell radius, and as above, A E is the
primary optics entry surface to the concentrator, and ηop is the optical efficiency.
The thermal resistance of the back plate working as heat distributor will be Rth =
ln(α Rc )/πwκ, which decreases with cell size. In general, for a fixed temperature
difference, the smaller the cell radius and the entry surface, the less the plate
thickness that will be required to maintain it. In the case of the back plate acting
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch11 page 530
as the only heat conductor and dissipater, its surface will be the same as that of
the entry surface of the concentrator and in this respect we could impose that
this surface is limited by the maximum effective radius as in Eq. (11.21), so that
A E = π Rc2 and Eq. (11.22) becomes
ln(α Rc )
Ti − To = −0.44B(n)ηop (11.23)
h eq
where κi is the thermal conductivity of the i th layer between cell and back plate
and di its thicknesses. However, for well-manufactured receivers most of the tem-
perature gap occurs in the back plate and the contribution of the upper layers is
small, so that their thermal properties can be to some extent disregarded. The use
of conductive glue instead of solder can create an exception.
A worked example with a real 500× module design allows us to see the
importance of the different elements of the encapsulation. Assume we are using
a triple-junction solar cell of area 0.5 cm2 , 150 µm thick, bonded with 50 µm of
solder to a copper tab 500 µm thick, which is then laminated to a thermal conductive
epoxy layer of 500 µm thickness, and then to 4 mm thick aluminium substrate slab
which is screwed to the module’s back plate, also made of aluminium and 2mm
thick. The surface of the Al slab is 16 cm2 and the Cu tabs and thermally conductive
epoxy printed in its surface cover 40% of this surface. Finally, in-between the Al
slab and the module’s back plate, thermal conductivity is enhanced by a 50 µm
thick layer of thermal grease. Taking h = 5 × 10−4 W cm−2 K−1 for the air–
back-plate interface and the emissivity of Al as 0.1, we find using Eq. (11.21)
that Rmax = 17.15 cm, which is greater than the radius of the 250 cm2 back-plate
dissipating surface, whether this is squared or circular. This means that all of is
being effectively used for heat dissipation. Assuming a direct normal irradiance
(DNI) value of 900 W m−2 , and an optical efficiency of 80%, we can calculate
the temperature difference between ambient temperature and the receiver bottom
using Eq. (11.22) as Ti − To = 56C.
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch11 page 531
Cell 0.67
Solder 0.36
Copper tab 0.03
Thermal epoxy 2.67
Aluminium slab 0.21
Thermal grease 1.87
Also using the respective thermal conductivities, thicknesses and areas of the
different layers, from the cell to the receiver’s bottom in contact with the back
plate, we would find the temperature jumps shown in Table 11.3.
The total thermal drop in the conducting layers adds up to 5.83◦C, which
added to the difference between the ambient temperature and the receiver makes a
total of 61.83◦ C, which is the temperature above ambient reached by the cell. This
example is good to show how most of thermal drop occurs in the module’s back
plate, performing as a heat sink. Care must be taken with the insulating layers —
the epoxy layer in our example — to ensure the required breakdown voltage that is
commonly specified in the 1–2 kV range is not exceeded. Here we have not taken
into account the convection in the Al substrate slab, which also reduces the overall
temperature drop somewhat.
In summary, cooling the cells is an important problem to tackle when facing
CPV module design, in which a combination of electric insulation and good thermal
contact has to be achieved. But in the end, it is helpful to know that simple passive
cooling means can provide the heat dissipation needed by the cells, even under
solar concentrations of 1000×.
Static mounts are only feasible today for low concentration factors below 5. In
the long term, static concentrators with higher ratios making use of luminescence
and photonic crystals might become available. However, all these issues are beyond
the scope of this work.
Line-focus reflective concentrators, such as troughs (the already-mentioned
mentioned EUCLIDES case), require only one-axis tracking to maintain the PV
receiver along the focus line. However, due to the daily variations in the Sun’s
elevation, the incidence of sunlight on the tracker’s aperture is usually somewhat
oblique, reducing the intercepted energy and causing the Sun’s image to move up
and down within the focal axis, producing further losses whenever it oversails the
receiver’s ends. Line-focus refractive concentrators (for example, those based on
linear Fresnel lenses, as in the ENTECH type) experience severe optical aberra-
tions when light incidence is not normal, thus requiring two-axis sun tracking, and
the same happens to most of the point-focus concentrators, excepting some low-
concentration factor devices with sufficient acceptance angle to admit the Sun’s
altitude variations.
Nearly all PV concentrators that are already commercialised or currently under
development use two-axis tracking, the so-called pedestal tracker, with its azimuth-
elevation axes, being the most common configuration, followed by the tilt-roll
tracker operating on the declination–hour angle axes shown in Fig. 11.24.
Strictly speaking, the main commitment to be fulfilled by a CPV Sun tracker is
permanently to align the pointing axis of the supported concentration system with
the local Sun vector, in this way producing maximum power output. The reasons
for the decrease of Sun-tracking performance can be classified into two main types
(Luque-Heredia et al., 2006): (1) those purely related to the precise pointing of the
Figure 11.24 The two most common Sun-tracking axes geometries used in solar trackers:
declination–hour angle (on the left) and azimuth-elevation tracking (on the right).
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch11 page 533
tracker at the Sun; and (2) those which provoke shrinkage of the overall acceptance
angle of the concentrator system, thus indirectly increasing the tracking accuracy
required. Among those related to the tracking accuracy are on the one hand the
exactness of the Sun’s positional coordinates generated by the control system,
expressed in terms of rotation angles of the tracking axes, either by Sun ephemeris-
based computations or derived from the feedback of Sun-pointing sensor readings,
or a combination of both, which are affected by numerous error sources. On the
other hand, there is the precision with which the tracker can be positioned at these
dictated orientations, i.e. the positioning resolution of the tracking drive and its
control system, which essentially depends on the performance of tracking speed
control and on the mechanical backlash introduced by the drive’s gearings.
Regarding acceptance angle losses caused by the tracking system, these are
due to the accuracy, which can be attained in the mounting and alignment of the
concentrator system atop of the tracker. This is basically a design problem having
to do with the fixtures provided for this purpose, their accurate assembly and
the means provided for in-field fine tuning, but also with the mounting protocols
devised to carry out these tasks. Also resulting in acceptance angle cuts is the
stiffness of the tracker, which is to say the deformations allowed in the different
elements of its structure under service conditions.
In a first iteration, characterisation of the service conditions for a CPV tracker
basically consist in determining the CPV array payload (modules’ weight) and fix-
ing the maximum wind load, i.e. wind speed, to be withstood during Sun-tracking
operation without any effect on the concentrator’s power output. To ensure service
under these maximum permanent and variable loads, a maximum flexure defor-
mation should be specified for the tracker structure.
As for the variable loads, the higher the maximum wind speed to be resisted
while maintaining productive operation, the heavier and more expensive is the
tracking structure required to keep deformation within the bounds required for
accurate tracking. This behaviour can be seen in Fig. 11.25 for one of Inspira’s
early tracking designs, a 9 m2 CPV pedestal tracker, in which normalised cost
vs. maximum service wind speed for a maximum 0.1◦ flexure is plotted (Luque-
Heredia et al., 2003).
A cost-effective approach here is to determine the maximum wind load from
the cross-correlation between wind speed and direct radiation, in the location or set
of locations in which the trackers are to be installed. A case example of this type of
analysis is presented in Fig. 11.26, worked out with one year of continuous wind
speed and direct normal irradiation hourly data (assuming a two-axis tracker), for
the Spanish city of Granada (Luque-Heredia et al., 2007). Above this threshold
stiffness specifications do not have to be met, service is not guaranteed, and the
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch11 page 534
Figure 11.25 Tracker cost vs. maximum service wind speed for a specific tracker design
dimensioned for three different aperture surfaces and 1000 units/yr production.
Figure 11.26 Wind speed vs. yearly DNI correlation for the determination of optimum
service conditions applied to Granada.
tracker would be better to switch to some low wind profile stow position to decrease
stress and increase operative lifetime. For example, as a general rule of thumb a
13 m s−1 maximum wind load both windward or leeward to the modules in any
of the Sun-tracking orientations of the tracker can be considered a reasonably
conservative value which has been proved to comprise a minimum of 95% of the
direct radiation measured by the 26 weather stations of the SOLMET network
distributed over the contiguous United States (Randall and Grandjean, 1982).
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch11 page 535
Regarding the CPV array payload, and again bearing in mind the targets set
for a possible CPV module design team in order to decrease tracker cost, it would
be worthwhile to decrease the module’s weight as much as possible; the weights of
present module technologies range from 40–140 g W−1. Moreover, module sizing
is also an important issue, and an optimum is to be sought in which the size of
modules does not require redundant framing from the tracker just to hold them,
but on the other hand in which module stiffness doesn’t develop into an excess of
self-weight.
The achievable tracking accuracy and the acceptance angle of the modules
determine the maximum deformation that can occur in the tracker structure without
affecting power output. The higher the tracking accuracy and the wider the nominal
acceptance angle of a module, the bigger the deformation tolerance which can be
set for the structure, and consequently the lower its cost.
Aside from the structural specifications derived from service conditions, basi-
cally determining the tracker’s stiffness, its ultimate limit state and the loads at
which this is reached is to be considered. This is usually a better-paved way in the
sense that there is a good set of international standard building codes. Even if these
do not yet consider the particular case of Sun-trackers, the CPV designer will be
able to interpret and adapt them, dimensioning structural ultimate resistance for
the recommended values of variable loads, namely wind and (depending on the
location) snow and earthquakes. Along with these static loads, dynamic effects
should also be taken into account in what respects the resonant frequencies of the
structure.
For a flat-plate tracker, acceptance angle is meaningless and thus ultimate limit
state dimensioning constitutes only structural specification; stiffness is only an
issue if excess bending can somehow stress or damage the PV modules. However,
in the case of CPV trackers it can easily happen that the stiffness specifications are
more stringent than the ultimate resistance ones, so the fulfilment of the former
implies that the latter are overridden. For this reason a CPV tracker, whichever
its tracking axes configuration, is frequently heavier than its conventional flat-
plate counterpart, and therefore somewhat more expensive. But on the other hand,
thanks to the fact that HCPV modules are already more than double the average PV
efficiency, tracker costs per unit Wp have the potential to outperform their flat-plate
counterparts.
Further exploring the sensitivity of tracker cost to flexure specifications,
Fig. 11.27 shows how for a 50 m2 CPV pedestal tracker designed by Inspira, weight
diminishes when maximum allowed flexure under maximum service conditions is
increased, apparently following a potential law, to the point that below 0.4◦ we
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch11 page 536
Figure 11.27 Steel weight vs. maximum flexure under maximum service conditions for a
50 m2 pedestal CPV tracker design.
Among these error sources the most significant have a deterministic nature
and mainly result from self-characterisation inaccuracies, i.e. tolerance deviations
appearing in the controller’s default assumptions regarding the tracker’s installa-
tion or manufacturing. Drifts in the internal timing required for the computation of
the Sun ephemeris are the other critical error source, which should be restrained.
Second-order error sources, such as gravitational bending in wide-aperture track-
ers, the effect of mismatch in multi-secondary axis trackers, or even ephemeris
inaccuracies due to the effect of local atmospheric refraction, might have to be
considered as well.
To suppress the tracking errors caused by the above-mentioned sources, feed-
back has to be integrated in the control strategy. This open-loop core strategy
blended with the fed-back closed-loop is sometimes referred to as the hybrid
approach, which in the past has been mainly connected with control theory (Luque-
Heredia et al., 2005a). Related to this approach but relying on the principles of
automated instrumentation calibration, and working in much the same way as the
pointing control of big telescopes, Inspira developed the SunDog sun-tracking
control unit (STCU) (Luque-Heredia et al., 2005b), whose operation is based on
the internal calculation of high-precision solar ephemeris (mean error ≈0.01◦ ).
To convert the solar coordinates generated by the ephemeris into tracking axes
rotation angles with respect to a specific reference orientation, the SunDog STCU
contains a non-linear calibration model characterised by a series of parameters that
determine the transformation in a precise manner. Those parameters are related to:
(1) the relative position of the tracking axes with respect to the Earth’s local sur-
face; (2) the orientation of the axes of rotation references; and (3) the orientation
of the pointing vector of the concentrator system with respect to the tracking axes.
The parameters of the model are least squares fitted internally by means of global
optimisation routines and from precise measurements of the Sun’s position, deter-
mined by maximising the concentrator’s power output. From then on the STCU
operates as an open-loop controller, with proven long-term accuracies of better
than 0.1◦ with 97% probability and below 0.05◦ with 60% probability, as shown in
Fig. 11.28. This was measured with Inspira’s SunSpear tracking accuracy sensor,
which is based on solid-state image sensors (Luque-Heredia et al., 2005c). This
tracking control technique was first used, in a different context, by Arno Penzias,
the discoverer of cosmic background radiation, for pointing early satellite receivers
by fitting an error model with position measurements of a precisely located radio
galaxy.
Also of the utmost importance to achieve Sun-tracking accuracy is the posi-
tioning resolution, i.e. the capacity to position the tracker at whatever command
position is produced by the controller, for example in the case of a SunDog STCU
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch11 page 538
Figure 11.28 Example of daily Sun-tracking error statistics, probability density and dis-
tribution function, with the probability density of collimated sun beam over the tracking
accuracy sensor surface superimposed, showing the 0.1◦ and 0.2◦ tracking error rings.
the position resulting from the Sun ephemeris computation and the later conver-
sion of these Sun coordinates into tracking axes rotation angles by the calibration
model. These high resolutions can usually be achieved by a mixture of versatile
speed control and low backlash gearings, which is in the case of step-based tracking
control schemes will permit a tracking step that is as small as possible, which in
the last instance will serve to decrease tracking error variance.
industry. Different corporations were created in the United States in the mid-1980s
around this concept, out of which, as we will see below, Amonix and Entech Solar
are still operating in 2013. Also, some important demonstration projects were
built, such as the above-mentioned EUCLIDES plant in Tenerife, which was until
2008 the biggest CPV plant in the world. However, concentration photovoltaic
projects were hardly ever developed further than their prototype stage, even when
they showed promising results in the laboratory, either because they did not reach
a commercialisation stage, or if they did, they could not develop the minimum
economies of scale necessary to demonstrate cost reductions.
The situation has changed radically since those early days of CPV. Because
of the booming growth experienced in the period 2000–2010, the photovoltaics
industry reached sufficient critical mass to boost business ideas based on alternative
technologies such as concentration.
With this fertile scenario, the rapid increase of multijunction cell efficiency
in recent years has played an important role to recall and reassert the promise
of the high cost-reduction potential of concentration systems. Likewise, the fact
that theoretically there is still plenty of room for efficiency increase by means
of increasing the number of junctions, and also in the medium term through the
introduction of other high- efficiency concepts now being researched, has taken
concentration systems, formerly based on more or less customised silicon cells
conceived to operate in the 20–100× range, to the so-called HCPV range. Here the
concentration factors range from 500× to 1000×, which is the level at which these
new high-efficiency cells can show a clear economic advantage. Using a simple
example, at a concentration ratio of 1000, a triple-junction cell cost of 100,000 $/m2
falls to 100 $/m2 , which at efficiency levels of 33% and normal direct irradiation
of 1000 W/m2 , results in cell costs of 0.3 $/W, as opposed to the 2.3 $/W cost of
conventional PV modules in 2011 (here, it must be noted, including also housing
costs).
Concentrated photovoltaics has evolved to become the platform from which
to launch high-efficiency cells onto the market, even though this market entry has
significantly slowed down due to the drastic price reductions in flat-plate silicon
modules that occurred in 2012/13. However, the tendency to achieve higher con-
centration ratios has been confronted with important technical challenges. As we
have seen, the higher the concentration factor, the smaller is the maximum possi-
ble acceptance angle of the optics. At a system level, additional acceptance angle
losses will also have to be subtracted from the nominal acceptance angles of sin-
gle concentrator units, these deriving from misalignments in the mounting of such
sub-systems in modules and subsequently of these on the tracking systems, and
also from structural flexure in the tracking system due to its payload or service
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch11 page 540
wind load. In spite of these difficulties, moving to high concentration factors also
yields some advantages. For example, moving to the use of small cells of at most
1 cm2 size helps to make heat dissipation easier and at some point also facilitates
the adoption of the manufacturing processes used by optoelectronic industries that
have achieved very high learning rates, i.e. extent of cost reductions as cumulative
production increases.
In sum, a quickly growing PV industry, the potential foreseen in high-efficiency
cells, and the knowledge gathered from past CPV prototypes and projects, all
together favoured the development of concentration photovoltaics until the eco-
nomic downturn of 2012–2013. And as a clear signal in this respect, more than
30 new CPV start-ups appeared during the first decade of the twenty-first century.
In Table 11.4 we give a list of some of these start-ups, choosing mostly those that
started some industrial production and have operating installations of a relevant
size, or those that represent an example of a novel design.
Guascor Fotón Point, Si, Fresnel 400× 12, 550 Subsidiary of engine
(Spain) (TP-R), 2 axis, manufacturer Guascor
Az&El, pedestal Redesign underway;
moving to higher
concentrations and
III–V, MJ cells
Renamed as Foton HC
Systems
Isofotón (Spain) Point, III–V, TIR-R 1000× 400 Business division within
(specific secondary), flat-plate PV
2 axis, Az&El, manufacturer Isofotón
pedestal Redesign underway;
moving to Fresnel +
secondary and lower
concentration
Opel (USA) Point, III–V, Fresnel 500× 470 Sold solar assets to
(TP-TIR), 2-axis, Northern States Metals
Az&El, pedestal in Dec. 2012
Semprius (USA) Point, III–V, Plano 1110× 8 Siemens acquires
convex (glass ball), minority share
2-axis, Az&El,
pedestal
Sharp (Japan) Point, III–V, Fresnel 700× 100
(−), 2-axis, Az&El,
pedestal
Skyline (USA) Linear, III–V, Mirrored 10× 27
Trough, 1-axis,
horizontal, NS
Silicon CPV Point, Si, Parquet of 120× 75 Also developing a III–V,
(UK) prisms, 2-axis, MJ 960× system
Az&El, pedestal
Sol3g (Spain) Point, III–V, Fresnel 476× 1000 Acquired by engineering
(TP-TIR), 2-axis, and construction
Az&El, pedestal company Abengoa
(Spain)
Redesign underway
Solar Systems Point, III–V, Mirrored 500× 1, 200 Acquired by engineering
(Australia) Parabolic dish, company Silex Systems
2-axis, Az&El, (Australia)
pedestal
(Continued)
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch11 page 543
a Focus geometry (linear, point); cell material (Si, III–V compound); type of primary optics; type
of secondary optics (TP: Truncated pyramid, TIR: Total internal reflection, R: Refractive); one- or
two-axis tracking; tracking geometry.
b Installed capacity in kW.
*This information was correct at the time of compilation (November 2011).
Figure 11.29 One of the ISFOC CPV plants at Puertollano, in which three different tech-
nologies can be seen. In the foreground is Soitec (formerly Concentrix), in the background
SolFocus and on the right is Isofotón. Courtesy of ISFOC.
(Rubio et al., 2010a; Gil et al., 2010; Alamillo et al., 2010). Some of these results
will be reviewed in the next section.
One of CPV’s problems has been the wide variety of different designs pro-
posed, as usually happens with any nascent industry (recall the early days of avia-
tion, for example). This lack of standardisation hindered the development of a set
of specific CPV component providers (e.g. trackers, optics, etc.) that could attain
cost reductions through aggregated market volume. However, in recent times, as
can be inferred from the different system descriptions in Table 11.4, CPV system
designs are converging towards a set of common features, firstly confirming the
general trend towards higher concentration ratios, which is in most cases achieved
through the use of Fresnel primary optics plus some kind of secondary, frequently
a simple reflective truncated pyramid. Regarding the tracker, the two-axis azimuth-
elevation pedestal configuration is the most-chosen option, unless when targeting
the rooftop market, in which case the tilt-roll tracker has the advantage of a lower
profile. Without debating the pros and cons of this coalescing standardisation, it did
indeed help the appearance of dedicated suppliers to the CPV industry, whether
start-ups or already-established companies creating product lines specifically to
supply CPV manufacturers. Also a symptom of this standardisation is that design
debates have become more specific. For example, given that a Fresnel primary is
usually chosen, the question now is should it be made in PPMA acrylic plastic
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch11 page 545
Figure 11.30 Built in 2008 by Spanish developer Parques Solares de Navarra, the Vil-
lafranca plant with 10 MW of Guascor Fotón CPV systems was the largest CPV installation
in the world until 2012, when the 30 MW Alamosa Solar Plant opened in Colorado. In the
upper-left corner of the picture, the systems in a darker tone are 2 MW of flat PV mounted
on two-axis trackers. Courtesy of Parques Solares de Navarra, SL.
Figure 11.31 Victor Valley College 1 MW plant built by Solfocus in 2010 was the first one
set by this company in the United States. Courtesy of Solfocus, Inc.
1 Sadly, SolFocus later fell victim to the fall in price of flat-panel silicon PV and in September 2013
was reportedly in negotiations to sell its assets and intellectual property (IP).
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch11 page 548
operation conditions — they are overly optimistic with respect to what is normally
found in real operation — but there is by now sufficient track record for PV plant
planners to know what the energy output will be for an installation rated under
standard conditions in a location with a known irradiance and temperature. There
are also well-accepted formulae that have been developed to transform electrical
output readings — current, voltage and power — from a PV module, made at what-
ever irradiance and cell temperature conditions, into those that would correspond
to STC.
A set of STC must also be established to rate CPV modules or systems. Having
those for flat PV modules as the market’s well-established reference, it seemed
reasonable to lower the irradiance value somewhat, given that CPV modules do not
capture diffuse irradiance as tracking flat-plate PV modules would do. However,
ultimately it is the energy production which is of most interest, and in order to
remain close to the way things are in the flat-plate PV market, it seems a good
approach to determine CPV’s STC as those that will result in a similar energy output
to that of an equivalent non-concentrating PV system, i.e. a two-axis tracking PV
system. This was conceptually the approach taken by Luque in the specifications for
ISFOC’s first call for tenders in 2007, where he estimated the STC to be 850 W m−2
and 60◦ C cell temperature. Obviously this estimate is site-specific, mostly through
the available DNI irradiance and the ambient temperature.
This approach was further analysed by ISFOC (Rubio et al., 2010b), who
showed that at Puertollano — the location of most of ISFOC’s installed CPV
power — STC defined as 850 W m−2 DNI and 60◦ C cell temperature resulted in
yearly cumulative production (kWh/kW) of CPV which was about 5% higher than
that of tracked PV in the same location. On the other hand, were we to choose
STC similar to those of flat-plate PV, i.e. 1000 W m−2 DNI and again 60◦ C cell
temperature, then the yearly cumulative CPV production would be almost 15%
below that of tracked PV. The best match in ISFOC’s analysis, in-between the
cumulative productions of CPV and tracked PV, when varying the STC DNI, is
obtained with 900 W m−2 DNI and 60◦ C cell temperature. Also a good match,
better than 5%, occurs when varying cell temperature instead, for example using
850 W m−2 and 25◦ C cell temperature. While the rationale behind this analysis
and approach to a definition of the CPV rating STC puts tracked PV and CPV
on the same level in energy production terms, some CPV manufacturers have
justifiably opted to raise DNI and lower cell temperature under which they rate
their products, while there is still no generally accepted STC, in order to somewhat
increase the nominal power of their products and therefore lower their /W, which
as has been said is still the market’s main metric. The obvious trade-off is that going
to more favourable STC would lower their kWh/kWp, eventually below those of
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch11 page 549
tracked PV systems, but initial sales experiences showed this was less important in
a marketing strategy. The issue of the appropriate STC for CPV has been debated
by the International Electrotechnical Commission (IEC) when working on its IEC
62670 standard. Although a good case was made by the IEC TC82 Working Group
7 for 900 W m−2 and 25◦ C cell temperature to be the standard, the IEC has recently
ruled that the STC for CPV will remain the same as those for flat-plate PV as regards
irradiance and cell temperature (1000 W m−2 and 25◦ C) but that the standard solar
spectrum will be AM 1.5 DNI rather than AM 1.5 G (Solar Industry, 2013).
Figure 11.32 On the left is a drawing of the complete solar simulator developed by IES-
UPM for the rating of CPV modules, providing an extended source of collimated light (with
a subtended angle equivalent to the Sun’s), through the reflection of the light of a Xenon
flash lamp in a 2 m accurate parabolic mirror, shown in the picture on the right. Courtesy
of Soldaduras Avanzadas, SL.
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch11 page 550
As is well known, one of the properties of a parabola is that every ray coming
from its focus is reflected back parallel to its axis, so that by correctly sizing the
diameter of the light source located at this focal point, the light reflected by the
mirror will have a uniform angular spread. For example, in the currently available
setup, a 2 m diameter parabolic mirror with a 6 m focal distance, at which a 6 cm
diameter bulb is placed, will ideally provide a ±0.275◦ spread beam in the area
covered by the parabolic mirror. The simulator operates as a multi-flash testing
system. The Xenon lamp is triggered to illuminate the CPV module that is biased
at a different voltage at each flash pulse, recording different current–voltage pairs
during the decay of the flash pulse. Each current measurement within a single
flash corresponds to a certain irradiance level, which can be determined through a
calibrated cell, so that at the end of the series of flashes, a family of I –V curves is
obtained at varying concentration factors. A computer controls the whole process
by triggering the lamp, setting the biasing voltage for each flash, and recording the
current–voltage pairs.
One of the main difficulties in developing this type of solar simulator was
to find a 2 m diameter parabolic dish, which due to the long focal distance
required demanded extremely accurate optical quality. This could only be found
at sky-high prices in the astronomical telescope industry. A parabolic profile was
therefore machined out of a bulk aluminium piece, and its mirroring provided by
an aluminium-based reflective coating. Errors in the profile are in the 10 µm range,
which is higher than the tolerance permitted in astronomical telescopes but good
enough for the CPV application. Scattering due to the mirror surface finishing has
the effect of widening the angular spread of the collimated beam up to ±0.43◦ ,
i.e. the angular size of the Sun plus some circumsolar radiation, which is still good
enough for most CPV modules.
The spectrum of the flash lamp varies over time during its discharge. This
can be monitored by means of isotypes, which are triple-junction solar cells with
a single cell electrically contacted. The result is that only one of the curves in
the family described above has the desired spectrum. With faster electronics it is
possible to draw the whole I –V curve in a single flash. Additional software can
provide spectrum-corrected I –V curves at different irradiances.
We will also describe a procedure devised for the power rating of complete
CPV plants, which is the one used by ISFOC in the acceptance tests it carries out
on its contractors’ plants and which constitute a contractual payment milestone
(ISFOC, 2006). First, the plant’s DC power is measured at the ISFOC-specified
STC (850 W m−2 , 60◦ C cell temperature) by tracing a set of I –V curves, using a
capacitive electronic load. This load essentially consists of a discharged capacitor
that is connected to the CPV array, so that it will start to charge itself as it receives the
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch11 page 551
array’s current, thus gradually increasing its voltage and decreasing the current,
sweeping in this way the I –V curve and recording a set of current and voltage
pairs. This type of electronic load was specifically developed at IES-UPM for the
measurement of high-power PV systems. Full 100 kW plant units can be measured
with it, in a rather inexpensive way (Antón, 2004).
At least 15 I –V curves are required by the ISFOC procedure: this repetition is
very important because the tracker is always slightly out of aim and the I –V curve
may vary by about 5% depending on the moment of the measurement. These must
be measured with DNIs over 700 W m−2 and wind speed values below 12 km/h.
Before and after tracing each of these I –V curves, the temperature in the back plate
of a few modules in the plant is measured and the average is recorded, together
with the DNI. Each of these I –V curves will be translated from its registered mea-
surement conditions of average back-plate temperature and DNI to STC, so that
the STC DC power rating of the plant, PDC , is the average of the maximum power
values of all I –V translated curves. The translation equations used by the ISFOC
procedure first calculate the cell temperature through the average temperature mea-
sured in the module’s back plates. A thermal resistance parameter ρT is introduced,
which is defined as dependent on the direct irradiance received by these modules:
Tcell − Tbackplate
ρT = (11.25)
B
where Tcell and Tbackplate are the cell and back-plate temperatures respectively, and
B is the measured DNI. Translation of current–voltage pairs from measurement
conditions to STC is obtained through
BSTC
ISTC = I
B
0.0257(TSTC − Tcell ) (I L1 − I )(I L2 − I )(I L3 − I )
VSTC =V +N ln
297 I L1 I L2 I L3
TSTC
+ [N(E g1 + E g2 + E g3 ) − VOC ] 1 − (11.26)
Tcell
where I, V are respectively the measured current and voltage, ISTC , VSTC the
current and voltage measured at STC, BSTC, TSTC the values for STC DNI and
cell temperature (all temperatures in Kelvin), and N is the number of array cells
connected in series. E g1 , E g2 , E g3 are the energy bandgaps, in eV, of each three
junctions in the cell, and I L1 , I L2 , I L3 the short-circuit currents of each three cell
junctions under the measurement conditions. When going through this rating pro-
cedure, ISFOC contractors could themselves provide the currents of the sub-cells
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch11 page 552
Figure 11.33 Comparison between cumulative energy of CPV, and tracked and fixed-mount
conventional (non-concentrating) PV systems monitored by ISFOC for 1.5 years. In this
case the CPV rating, 850 W m−2 and 25◦ C cell temperature, produces an energy output
very similar to that of tracked PV and as can also be seen the cumulative DNI evolves very
closely to the cumulative production, implying a performance ratio value for the case of
CPV very close to unity. Reprinted with permission from Rubio et al. (2010b), © AIP.
above was the STC best option found by ISFOC. It is worth noting that this is the
overall production of all the CPV plants at Puertollano, regardless of differences
in their technologies, and without filtering any outage periods due to research and
development (R&D) works, maintenance, etc.
It is also interesting to check energy production on a monthly basis. We can
see in Fig. 11.34 that energy production in Puertollano during the summer months
with the highest DNI is up to 25% better than that of tracked PV and as much as
65% better than fixed-mount PV. The good performance of CPV plants during the
high DNI months is due not only to intrinsic module efficiency increase, but also
to the more fluent operation of the inverters, which under clear skies do not suffer
from intermittent connections and disconnections due to passing clouds. This can
be seen more clearly in Fig. 11.35, where efficiency of CPV and tracked PV plants
are plotted as a function of the DNI: while CPV’s performance increases linearly
with DNI, silicon PV’s efficiency linearly worsens due to its higher temperature
coefficient.
The inverter can also be a source of efficiency losses in the high DNI months
if it is undersized, as can be seen in ISFOC’s analysis of one of its concentrator
systems that had a 6 kW inverter per tracker, which was found to be undersized
during the summer, thus capping system power and therefore efficiency. Thinking
engineering-wise this undersizing should be avoided; however, the sizing of the
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch11 page 554
Figure 11.34 Monthly cumulative energy of the same three types of installations as in
Fig. 11.33, showing how CPV outperforms tracked PV in the summer months. Reprinted
with permission from Rubio et al. (2010b), © AIP.
Figure 11.35 Monthly cumulative energy of the same three types of installations as in
Fig. 11.34, showing how CPV outperforms tracked PV in the summer months. Reprinted
with permission from Rubio et al. (2010b), © AIP.
inverter can get to be an important and rather subtle issue in a solar developer’s
financial spreadsheet, as certain market situations have proved to favour the so-
called repowering of PV plants, i.e. inverters being undersized by as much as 30%
(Franco de Saravia and Rossel, 2010). In CPV it has been a common practice
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch11 page 555
from the beginning to install one inverter per concentrator system of power in
the 5–10 kW range. However, there is no special advantage in doing this (Alamillo
et al., 2010) and it is a fact that centralised inverters — i.e. a single inverter servicing
several concentrator systems of up to 500 kW — are as much as 35% cheaper than
the small 5 kW ones (Knoll, 2010). Some inverter manufacturers have reported that
power vs. operating voltage of concentrator fields, as these grow bigger, present
several maxima, thus requiring MPPT (maximum power point tracking) algorithms
that seek the global maximum. Usual P&O (peek & observe) algorithms will not
then be good enough to lock the global maximum, and voltage sweeps should be
programmed periodically.
Also among the ISFOC’s first reports on its CPV plants were those dealing
with reliability issues (Gil et al., 2010). Classification and statistics of defects,
detected during plant installation, prior to its commissioning and acceptance by
ISFOC, were undertaken for 2 MW of total installed power, distributed in 20 plants
of 100 kW, comprising five different CPV technologies. The mechanical tracker
was reported as the most flaw-prone during commissioning, as can be seen in
Table 11.5.
Note that the mechanical tracker is considered separately from the tracking
controller. As regards the mechanical tracker, ISFOC’s report also distinguishes
between the tracker’s steel structure, the tracking drive, and the foundation, finding
that 76% of mechanical tracker problems correspond to the structure and 23%
to the drive. Problems detected in the tracker structure are mainly due to faulty
transportation, which has either bent the steel profiles or damaged their coating,
with rust points appearing. Concentrated photovoltaics tracker transportation to the
field and assembly and leveling of modules on site are possibly areas with much
room for improvement. Regarding the drive problems, most of these seemed to be
localised in the elevation axis, usually addressed by all tracker designs using power
jacks. Regarding the second most important source of problems, in the wiring or
electrical protection, these were usually related to defective layout and overlooking
electrical safety codes.
In 2011, ISFOC reported in respect of degradation and O&M issues 2.5 years
after the installation of its oldest plants (Rubio et al., 2011). After a threefold
analysis checking for efficiency reductions, they found no evidence of noticeable
degradation having happened during this time.
To end, we have referred to the work being done towards the CPV rating norm
within IEC, and it must also be mentioned that the first CPV-specific standard
already approved and published is IEC 62108 (IEC, 2007). Essentially this is a set
of accelerated aging tests devised to detect long-term reliability problems in the
CPV modules (Muñoz et al., 2010).
Table 11.6 HCPV system cost forecast by Yamaguchi and Luque (1999), reprinted with
permission, © IEEE.
Short-term Long-term
1Ja 2J No learning Learning
Substrate wafer ($ per cm2 wafer area) 8.50 8.50 8.50 2.37
Cells ($/cm−2 cell area) 13.4 15.85 15.85 4.43
Cells ($/m−2 module area) 134 159 159 44
Optics, heat sink & assembling ($/m−2 ) 131 131 131 69
Module ($/m−2 ) 265 290 290 113
Cell efficiency (%) 23.1 37 45 45
Module efficiency (%) 19.0 30.5 37.1 37.1
Module ($/W) 1.39 0.95 0.78 0.31
Area-related BOS ($ per m2 aperture 114 114 114 60
area)
Power-related BOS ($/W) 0.22 0.22 0.22 0.12
Plant price ($ per m2 aperture area) 526 589 607 271
Madrid DNI (kWh m−2 year−1 ) 1826 1826 1826 1826
Performance ratio∗ 0.606 0.606 0.606 0.606
Electricity costs in Madrid ($/kWh) 0.186 0.130 0.110 0.050
Electricity costs in EGLb ($/kWh) 0.131 0.091 0.077 0.035
which it was derived is clearer than that of present-day forecasts by market research
firms, having at least a didactic value.
The cost analysis was produced for 1000× concentration modules that used
III–V cells of 1 mm2 surface area and integrated different, highly compact optical
designs produced with the SMS method. This CPV technology was jointly devel-
oped by IES-UPM and Isofotón (Algora et al., 2000). The module design followed
the so-called ‘LED-like approach’ in that the assembly processes proposed resem-
bled those in the optoelectronics industry (Algora et al., 2006). The levelised cost
of electricity (LCOE) for an amortisation period of 30 years was calculated for the
average DNI of Madrid (1,826 kWh m−2 yr−1 ) and also for that of an extremely
good DNI location (EGL in Table 11.6 above) with 2,600 kWh m−2 yr−1 .
Single- and double-junction cells were the state of the art at the time (first and
second column in Table 11.6, with header 1J and 2J) and the authors, assuming
average prices in the wholesale electricity markets of 0.05 $/kWh, and customer
electricity prices of 0.1 $/kWh, concluded that CPV electricity could become cost-
effective for the customer at EGL locations. Overall system prices, including CPV
module and area-related BOS (balance of system) (which were calculated based
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch11 page 558
on the EUCLIDES plant data) envisaged for the 1J and 2J cell CPV systems, were
respectively 1.62 $/W and 1.17 $/W and were calculated for a cumulative produc-
tion of 100 MW, with a 25% mark-up applied on top to account for installation
costs and commercial profit margin. The learning rate, LR, of a certain product
or industry is the cost reduction achieved every time the cumulative production
doubles (Nemet, 2005), and can be calculated from
−b
qt
Ct = C 0 where LR = (1 − 2−b ) (11.28)
q0
with C0 being the known cost of a product at some past time and Ct the projected
cost at some future time, q0 and qt the cumulative productions achieved at these two
times, and b the so-called learning coefficient; the value b = 0.175, appropriate
for the PV industry of the time, was assumed.
The short-term projections when extended into the long term, supposing mod-
ules with four-junction tandem cells (third and fourth columns in Table 11.6) for
which module efficiencies of 37% could be expected, are so reasonably given
that present-day three-junction cells are now achieving module efficiencies in the
32–35% range. One of the authors of this chapter has written a paper in which his
belief that 50% cells will eventually be achieved is substantiated (Luque, 2011).
In the long-term case, Table 11.6 shows two options (third and fourth columns),
firstly with no learning, i.e. no consideration of cost reduction due to production
scaling, and secondly integrating the cost reduction achieved when arriving at a
cumulative production of 1 GW with a 32% learning rate, which is that exhibited
by electronics and optoelectronics. A higher learning rate than that achieved by
conventional PV at that point, was used for this long-term estimate, considering on
the one hand that CPV is much less mature than PV and therefore could achieve
faster cost reductions in its initial industrialisation, and taking the optoelectron-
ics reference as that closer to the LED-like design approach to HCPV modules.
This last long-term forecast, assuming learning, is capable of matching wholesale
electricity prices after a relatively low cumulative volume by today’s standards.
An important positive surprise when looking at this table is the value assumed
for the performance ratio (PR). This is defined as the ratio between the energy
produced and the rated power installed times the ‘equivalent hours’ of operation of
the modules. The latter are the energy falling into a unit module surface divided by
the rating solar irradiance for the modules. In the easy case that the rating irradiance
in the modules is 1 kW m−2 the equivalent hours are exactly the kWh of solar energy
incident on 1 m2 of concentrator aperture. In Table 11.6 this rating is used and the
PR value is very low (0.606) as deduced from that measured in the EUCLIDES
prototype. However, the data registered at ISFOC for three-junction solar cells are
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch11 page 559
much higher, 0.82 vs. 0.72 for the silicon flat module plants in the area. This better
behaviour is due to the smaller temperature effect in the three-junction solar cells
as compared with the silicon cells and this also explains the low value (0.606) for
concentrated silicon systems. Of course, the silicon tracking system collects global
normal irradiance (GNI) while the CPV system collects DNI, which in this site is
76% of the former. That is why the concentrator system produced only 87% of the
flat-panel’s output if the rating is the same for both (1 kW m−2 and 25◦ C) but the
ratio is not that of DNI/GNI but much better. With the measured PR (0.82) instead
of the one derived from Si cells experiments (0.606) the cost of CPV electricity is
substantially reduced.
Figure 11.36 CPV system cost estimated by Algora (2007) as a function of the concentra-
tion factor and the cell efficiency for two different scenarios of cumulative production.
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch11 page 560
for example the estimated final price for an installation operating at 400× is about
1 /W higher than if the installation operates at 100 Suns, almost independent of
the cell efficiency after 10 MW of cumulative production.
Also in this analysis, to further stress the cost-reduction leverage of increasing
the concentration factor, a system cost breakdown is provided for two different
cases. The pie chart on the right in Fig. 11.37 represents the contribution of the
different CPV plant cost elements, when using 40% three-junction cells operating
at 250× (after 10 MW of cumulative production), resulting in a final cost of 3.8
/W at plant level, mostly due to the big impact (66%) of the expensive solar cells,
while on the left cheaper 33% two-junction solar cells operating at 1000× achieve,
also after 10 MW of cumulative production, a much lower CPV plant cost of 2.4
/W. In this way, Algora (2008) shows the importance of concentration factor as
a CPV system cost reduction factor, in particular in the starting phase when mass
production costs have not yet been achieved.
the invoice issued by the PV manufacturer to the plant developer or the appointed
EPC firm — i.e. it is what you will pay for the PV system if you want to build a
plant, whether to operate it for its whole lifetime or to resell just after it is built.
The LCOE is defined as the cost of a unit of electricity in current monetary
units. It incorporates the /W of installed costs described above, and adds the
operation and maintenance costs over the life cycle of the plant or system, also
known as the total life-cycle costs (TLCC). These costs are divided by the energy
generated, which is a function of the DNI irradiation in the plant’s location, so that
the overall expression is
TLCC
LCOE = N Qn
(11.29)
n=1 (1+d)n
where Q n is the energy output in year n, N is the number of years of the analysis
period, usually the total number of productive years, and d is the discount rate used
to discount back future annual energy output to a present value. The discount rate
is the rate of return that could be earned on an investment in the financial markets
with similar risk.
The LCOE should be the most relevant metric for PV plant operators, for
example the so-called IPPs (Independent Power Producers) that enter into PPAs
(power purchase agreements) with electric utilities. This metric also has strong
advocates among present CPV manufacturers, who are usually elusive in public
regarding their installed costs ( /W), presumably because they yet do not bear
favourable comparison with those of conventional flat-plate PV, but frequently
present aggressive LCOE estimates and projections, and propose that LCOE is
standardised as the main commercial metric, instead of the /W of installed costs.
However, while it makes sense to base the financial assessment of investments
in PV installations on their LCOE values, in general the PV market does not yet
permit CPV manufacturers to base their marketing strategies on LCOE values, for
two main reasons. The first is that there is still no confidence in the O&M cost
figures and the long-term predicted power outputs of a new technology such as
CPV, but it is these figures that are used to compute the LCOE; this problem is
very much linked to that already mentioned: CPV’s lack of track record. In other
words, potential CPV customers frequently prefer to know about installed system
prices in order to easily compare with other PV technologies, and afterwards, if
they are in for a long-term investment, they try and make the LCOE calculation by
themselves with all sorts of safety margins.
The second reason is that frequently, where PV plants are financed by banks
through project finance, as happens in the FIT-based markets in Europe, there is
only a small set of strong EPC companies that are considered eligible to receive
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch11 page 562
funding (i.e. bankable) to build any plant. The EPC role will be to build a turnkey
plant and hand it over to the developer — the one who has received the fund-
ing — for its operation, and therefore EPC’s involvement in the plant ends after
two or three years, and EPC companies are not too much concerned about met-
rics such as LCOE to assess the long-term profitability of the plant; their main
target is to maximise profit in the plant construction (within quality margins)
and for this purpose it is /W of systems bought which is really of interest
to them.
To get a feeling of HCPV costs in recent years, it is helpful to refer to mar-
ket research done by specialised firms such as the above-mentioned CPV Today
(Extance and Márquez, 2010). In this report, the main HCPV industry players
(defining HCPV as geometric concentrations above 300×) were surveyed to deter-
mine current and projected installed costs, as well as the breakdown of costs among
system components. As can be seen in Table 11.7, the surveyed HCPV companies
were grouped into three categories, according to their installed costs, and averages
were given — to keep each company’s data anonymous — for each group. The
lowest installed cost group offered quite a low average, falling in the range of
flat PV modules. However, the authors noted that this group included a set of
companies ranging from recent HCPV entrants who claim new approaches which
can attain pronounced cost reductions (but which remain to be proven) to the most
experienced HCPV vendors, who may be attaining the production volumes and
experience necessary to bring costs down, but may be selling at a loss to enter the
PV market and develop volume.
Probably more realistic is to take the survey’s overall average of 3.7 /W,
which can be compared with the turnkey price of a tracked PV system which in
summer 2011 could be assumed as around 2.8 /W.
CPV Today also presented the averaged cost breakdown for HCPV systems,
as shown in Fig. 11.38a, derived from their survey. The assembly costs in the pie
Table 11.7 Costs of HCPV (over 300×) according to CPV Today’s market survey
(Extance and Márquez, 2010).
Projected installed
Installed cost 2010 cost 2015
Segment ( /W, arithmetic mean) ( /W, arithmetic mean)
Figure 11.38 Two different cost breakdowns for CPV. a) An average one produced by CPV
Today after polling a wide set of CPV manufacturers; b) that presented by Guascor Fotón
derived from their experience for a 5 MW CPV plant with their technology. Reprinted with
permission from Extance and Márquez (2010), © FC Business Intelligence. Courtesy of
Guascor Fotón, SA.
chart were those related to the module integrating cells and optics. Figure 11.38b
shows the CPV cost breakdown of Guascor’s technology (Aurtenetxe, 2010) for a
5 MW PV plant, this company being the only that had built plants of this size at
that time.
11.10 Conclusions
By the end of 2011, when this chapter was written, CPV technology had entered
a phase of incipient commercialisation. Dozens of companies were involved in
its development and the most promising of them, in our opinion, were the HCPV
technologies, fuelled by the fantastic efficiency achievements of MJ solar cells.
Most companies currently using high-efficiency silicon technology were moving
to adapt MJ cells to their products. The rapid fall in flat-plate silicon modules in
2012 and 2013 has affected the commercial development of CPV, as it has for some
of the other PV technologies described in this book.
It is predicted that the efficiency of MJ cells will continue to increase from their
present 44% towards the 50% efficiency to be achieved with five or six junctions
or perhaps with third-generation approaches, such as the intermediate-band solar
cell (Luque and Martí, 1997).
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch11 page 564
Trends towards high concentration are underway from the 500 Suns or less of
many present-day solutions towards the 1000 Suns or more to be expected in the
near future. This increase will have a very important benefit to the cost of CPV
systems in the short term, while the production volume of MJ cells is too small to
provoke a reduction-by-experience of the cell cost.
Concentrator optics is well understood today and the main challenges are
solved at least theoretically: good efficiency, large angular aperture and good illu-
mination homogeneity. Challenges to come are maintaining these properties with
the more demanding 1000× technology under way. In spite of these challenge,
improvements are to be expected that will lead to practical optical efficiencies of
90% as compared with today’s typical values of 80%.
Simple solutions suffice for heat management. Passive cooling without fins
is generally adopted and, despite the high concentration, over 1000 Suns, the cell
temperature is usually in the range of 80–90◦C, which high-performance cells can
easily cope with (Vázquez et al., 2007; Núñez et al., 2010).
Field experience is still small, but the ∼10 MW commercial plant in Vil-
lafranca, (Spain), the largest in the world today (still with Si cells), and the 30
MW Alamosa Solar Plant in Colorado (with MJ cells) show that the time of no
more than prototypes has gone. The establishment of the Institute for Photovoltaic
Concentration Systems (ISFOC) in Spain by 2006, and the 3 MW international
call-for-tenders launched there, have been important steps to help companies to
develop MJ cell technology. Module manufacturing has matured, and from the
∼20% efficiency at the module level by about 2008, when the plants for ISFOC
were delivered, we are already at over 30% module efficiency at the time of writing
(summer 2011). A module efficiency of 45% is foreseen (Luque, 2011).
Trackers are an important element of the systems cost, constituting maybe
a third of the total. Therefore this has been carefully studied and the conditions
for producing them at a cost per unit similar to those used for unconcentrated
arrays with tracking are clear: excellent control and mechanisms and wide-aperture
optics. Moreover, the high efficiency of MJ technology reduces by a factor of two
to three the cost per watt of the tracker as compared with that for unconcentrated
arrays.
Very good news is the high performance ratio of the MJ concentrator cells,
over 82% as compared to 72% for unconcentrated silicon cells and about 60% for
Si concentrated cells. This is attributable to the better temperature performance
of MJ solar cells. This may have a very important influence on the LCOE for
concentrators.
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch11 page 565
Ultimately CPV will develop if the prices are below those for unconcentrated
approaches. For this they must climb the barrier of the low-cost conventional uncon-
centrated photovoltaics, fuelled by 50 GW of manufacturing experience. Yet aca-
demic studies of CPV in the late 1990s predicted very low costs in the long term,
comparable or below most electricity sources, and by looking at these studies again
today, we think these goals can be achieved, if the above-mentioned barrier can
be overcome. If this is the case, CPV would become the leading technology for
large PV plants, which in its turn is the largest and fastest growing PV application
today.
References
Alamillo C., de la Rubia O., Gil E., Hipólito A., Martín A., Rubio F., Pachón J. L. and
Banda P. (2010), ‘Analysis of inverter configuration on CPV Plants’, Proc. 25th Euro-
pean Photovoltaic Solar Energy Conf., Valencia, pp. 4733–4736.
Alarcón J., Bassy A., Blazquez L., Luque A., Sala G. and Lorenzo E. (1982), ‘Central
fotovoltaica flotante’, Mundo Electrónico 117, 95–99.
Algora C. (2007), ‘Very-high-concentration challenges of III–V multijunction solar cells’,
in Luque A. and Andreev V., Concentrator Photovoltaics, Springer Verlag, Berlin,
pp. 89–112.
Algora C., Miñano J. C., Benítez P., Rey-Stolle I., Álvarez J. L., Díaz V., Hernández M.,
Ortiz E., Muñoz F., Peña R., Mohedano R., Luque A., Smekens G., de Villers T.,
Andreev V., Khostikov V., Rumiantsev V., Schvartz H., Nather H., Viehmann K. and
Saveliev S. (2000), ‘Ultra compact high flux GaAs cell photovoltaic concentrator’,
Proc. 16th European Photovoltaic Solar Energy Conf., Glasgow, pp. 2241–2244.
Algora C., Ortiz E., Díaz V., Peña R., Andreev V. M., Khvostikov V. P. and Rumyantsev R. D.
(2001), ‘A GaAs solar cell with an efficiency of 26.2% at 1000 suns and 25.0% at 2000
suns’, IEEE Trans. Electron Dev. 48, 840–844.
Algora C., Rey-Stolle I., Galiana B., González J. R., Baudrit M. and García I. (2006), ‘Strate-
gic options for a LED-like Approach in III–V concentrator photovoltaics’, Conf. Record
4th World Conf. Photovoltaic Energy Conversion, Waikoloa, Hawaii, pp. 741–744.
Algora C., Rey-Stolle I., García I., Galiana B., Baudrit M., Espinet P., Barrigón E., Datas
A., González J. R. and Bautista J. (2008), ‘A dual junction solar cell with an efficiency
of 32.6% at 1000× and 31.0% at 3000× ’, Proc. 5th Int. Conf. on Solar Concentrators,
Palm Desert, CA.
Andreev V., Ionova E., Rumyantsev V., Sadchikov N. and Shvarts M. (2003), ‘Concentrator
PV modules of ‘all-glass’ design with modified structure’, Conf. Record 3rd World
Conf. Photovoltaic Energy Conversion, Osaka, pp. 803–876.
Antón I (2004), ‘Métodos y equipos para la caracterización de sistemas fotovoltaicos de
concentración’, PhD Thesis, Universidad Politécnica de Madrid.
Araújo G. L. and Martí A. (1994), ‘Absolute limiting efficiencies for photovoltaic energy
conversion’, Solar Energy Mater. Solar Cells 90, 1068–1088.
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch11 page 566
Aurtenetxe A. (2010), ‘Current and future HCPV competitiveness in utility scale installa-
tions’, 3rd Concentrated Photovoltaics Summit, Sevilla, 18 November.
Benítez P., Miñano J. C., Zamora P., Mohedano R., Cvetkovic A., Buljan M., Chaves J. and
Hernández M. (2010), ‘High performance Fresnel-based photovoltaic concentrator’,
Optics Express 18, A25–A40.
Bruton T. M., Heasman K. C., Nagle. J. P., Cunningham D. W. and Mason N. B. (1994),
‘Large area high efficiency silicon solar cells made by laser grooved buried grid
process’, Proc. 12th European Photovoltaic Solar Energy Conf., Amsterdam. H. S.
Stephens & Associates, Bedford, pp. 761–765.
Burgess E. L. and Pritchard D. A. (1978), ‘Performance of a one kilowatt concentrator photo-
voltaic array utilizing active cooling’, Conf. Record 13th IEEE Photovoltaic Specialists
Conf., Washington D.C. IEEE Press, Piscataway, NJ, pp. 1121–1124.
Domínguez C., Antón I. and Sala G. (2008), ‘Solar simulator for concentrator photovoltaic
systems’, Optics Express 19, 14894–14901.
Extance A. and Márquez C. (2010), The Concentrated Photovoltaics Industry Report 2010,
FC Business Intelligence, London.
Franco de Saravia C. and Rossel A. D. (2010), ‘Power Boost’, Photon International 12,
44–52.
Friedman D. J., Olson J. M. and Kurtz S. (2011), ‘High efficiency III–V multijunction solar
cells’, in Luque A. and Hegedus S. (eds.), Handbook of Photovoltaic Science and
Engineering, John Wiley & Sons, Chichester, UK.
Garboushian V., Roubideaux D. and Yoon S. (1996), ‘An evaluation of integrated high-
concentration photovoltaics for large-scale grid connected applications’, Conf. Record
25th IEEE Photovoltaic Specialists Conf., Washington D.C. IEEE Press, Piscataway,
NJ, pp. 1373–1376.
Gil E., Sánchez D., de la Rubia O., Alamillo C., Rubio F. and Banda P. (2010). ‘Field
technical inspection of CPV power plants for the 25th EU PVSEC/ WCPEC-5’, Conf.
Record 25th European Photovoltaic Solar Energy Conf., Valencia, pp. 4483–4486.
Green M. A., Emery K., King D. L., Hisikawa Y. and Warta W. (2010), ‘Solar efficiency
tables (version 36)’, Progr. Photovoltaics 18, 346–352.
Guter W., Schöne J., Philipps S. P., Steiner M., Siefer G., Wekkeli A., Welser E., Oliva E.,
Bett A. W. and Dimroth F. (2009), ‘Current-matched triple-junction solar cell reach-
ing 41.1% conversion efficiency under concentrated sunlight’, Appl. Phys. Lett. 94,
223504.
IEC (2007), IEC 62108: 2007 Concentrator photovoltaic (CPV) modules and assemblies —
design qualification and type approval, International Electrotechnical Commission,
Geneva, Switzerland.
ISFOC (2006), ‘Specifications of general conditions for the call for tenders for con-
centration photovoltaic solar plants for the Institute of Concentration Photovoltaic
Systems’, Instituto de Sistemas Fotovoltaicos de Concentracion, Puertollano, Spain,
pp. 34–39.
King R. R., Boca A., Hong W., Liu X.-Q., Bhusari D., Larrabee D., Edmondson K. M.,
Law D. C., Fetzer C. M., Mesropian S. and Karam N. H. (2009), Proc. 13th European
Photovoltaic Solar Energy Conf., Hamburg, pp. 55–61.
Knoll B. (2010), ‘Still falling: inverter prices in the German spot market continue to get
cheaper’, Photon International 12, 108–109.
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch11 page 567
Kolodinski S., Werner J. H., Wittchen T. and Queisser H. J. (1993), ‘Quantum efficiencies
exceeding unity due to impact ionization in silicon solar cells’, Appl. Phys. Lett. 63,
2405–2407.
Landsberg P. T. and Tonge G. (1980), ‘Thermodynamic energy conversion efficiencies’,
J. Appl. Phys. 51, R1–R20.
Lewis N. S., Crabtree G., Nozik A. J., Wasielewski M. R. and Alivisatos P. (2005), ‘Basic
research needs for solar energy utilization’, Office of Science — US Department of
Energy, http://science.energy.gov/∼/media/bes/pdf/reports/files/seu_rpt.pdf.
Leutz R. and Suzuki A. (2001), Non Imaging Fresnel Lenses, Springer Verlag,
Berlin/Heidelberg.
Lorenzo E. and Sala G. (1979), ‘Hybrid silicone–glass Fresnel lens as a concentrator for
photovoltaic applications’, in Sun II: Proc. Silver Jubilee Congress, Pergamon Press,
Atlanta, GA, pp. 536–539.
Lorenzo E. and Luque A. (1981) ‘Fresnel lens analysis for solar energy applications’, Appl.
Optics 20, 2941–2945.
Luque A. (1980), ‘Quasi-optimum pseudo-lambertian reflecting concentrator: an analysis’,
Appl. Optics 19, 2398–2402.
Luque A. (1986), ‘Non-imaging optics in photovoltaic concentration’, Phys. in Technology
17, 118–124.
Luque A. (1989, ed.), Solar Cells and Optics for Photovoltaic Concentration, Adam Hilger,
Bristol.
Luque A. (2011), ‘Will we exceed 50% efficiency in photovoltaics?’, J. Appl. Phys. 110,
031301–031301-19.
Luque A., Gidon P., Pirot M., Anton I., Caballero L. J., Tobias I., Canizo C. del and
Jausseaud C. (2004a), ‘Performance of front contact silicon solar cells under con-
centration’, Progr. Photovoltaics 12, 517–528.
Luque A. and Martí A. (1997), ‘Increasing the efficiency of ideal solar cells by photon
induced transitions at intermediate levels’, Phys. Rev. Lett. 78, 5014–5017.
Luque A. and Martí A. (2011), ‘Theoretical limits of photovoltaic conversion and new
generation solar cell’, in Luque A. and Hegedus S. (eds.), Handbook of Photovoltaic
Science and Engineering, John Wiley & Sons, Chichester, UK.
Luque A. and Martí A. (2012), ‘Understanding intermediate-band solar cells’, Nature Pho-
tonics 6, 146–152.
Luque A., Sala G., Arboiro J. C., Bruton T., Cunningham D. and Mason N. (1997), ‘Some
results of the EUCLIDES photovoltaic concentrator prototype’, Progr. Photovoltaics
5, 195–212.
Luque A., Tobias I., Gidon P., Pirot M., Canizo C. del, Anton I. and Jausseaud C. (2004b),
‘Two-dimensional modeling of front contact silicon solar cells’, Progr. Photovoltaics
12, 503–516.
Luque-Heredia I., Cervantes R. and Quéméré G. (2005c), ‘A sun tracking error monitor
for photovoltaic concentrators’, Conf. Record 4th World Conf. on Photovoltaic Energy
Conversion, Hawaii, 706–709.
Luque-Heredia I., Martín C., Mañanes M. T., Moreno J. M., Auger J. L., Bodin V.,
Alonso J., Díaz V. and Sala G. (2003), ‘A subdegree precision sun tracker for 1000×
micro-concentrator modules’, Conf. Record 3rd World Conf. on Photovoltaic Energy
Conversion, Osaka, Japan, pp. 857–860.
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch11 page 568
Luque-Heredia I., Moreno J. M., Magalhães P. H., Cervantes R., Quéméré G. and Laurent O.
(2007), ‘Inspira’s CPV sun tracking’, in Luque A. and Andreev V. (eds), Concentrator
Photovoltaics, Springer Verlag, Berlin, pp. 221–252.
Luque-Heredia I., Moreno J. M., Quéméré G., Cervantes R. and Magalhães P. H. (2005a),
‘CPV Sun tracking at Inspira’, Conf. Record 3rd Int. Conf. on Solar Electric Concen-
trators for the Production of Electricity or Hydrogen, Scottsdale, AZ.
Luque-Heredia I., Moreno J. M., Quéméré G., Cervantes R. and Magalhães P. H. (2005b),
‘SunDog STCU TM generic sun tracking unit for concentration technologies’, Conf.
Record 20th European Photovoltaic Solar Energy Conf., Barcelona, pp. 2047–2050.
Luque-Heredia I., Quéméré G., Magalhães P. H., Fraile de Lerma A., Hermanns L., de
Alarcón E. and Luque A. (2006), ‘Modelling structural flexure effects in CPV sun
trackers’, Conf. Record 21st European Photovoltaic Solar Energy Conf. and Exhibition,
Dresden, Germany, pp. 2105–2110.
Miñano J. C. (1985a), ‘Two-dimensional non-imaging concentrators with inhomogeneous
media: a new look’, J. Opt. Soc. Amer. A 2, 1826–1831.
Miñano J. C. (1985b), ‘Design of three-dimensional optical concentrators with inhomoge-
neous media’, J. Opt. Soc. Amer. A 3, 1345–1353.
Miñano J. C. (1985c), ‘New family of 2-D non imaging concentrators: The compound
triangular concentrators’, Appl. Optics 24, 3872–3876.
Miñano J. C. and González J. C. (1992), ‘New method of designing of non-imaging con-
centrators’, Appl. Optics 31, 3051–3060.
Miñano J. C., González J. C. and Benítez P. (1995), ‘RXI: a high gain, compact, non-imaging
concentrator’, Appl. Optics 34, 7850–7856.
Muñoz E., Vidal P. G., Nofuentes G., Hontoria L., Pérez P., Terrados J., Almonacid G. and
Aguilera J. (2010), ‘CPV standardization: an overview’, Renewable and Sustainable
Energy Rev. 14, 518–523.
Napoli L. S., Swartz G. A., Liu S. G., Klein N., Fairbanks D. and Tamatus D. (1977), ‘High
level concentration of sunlight on silicon solar cells’, RCA Review 38, 76–108.
Nemet G. F. (2005), ‘Beyond the learning curve: factors influencing cost reductions in
photovoltaics’, Energy Policy 34, 3218–3232.
Ning X., Winston R. and O’Gallagher J. (1987), ‘Dielectric totally internally reflecting
concentrators’, Appl. Optics 26, 300–305.
Núñez N., Vázquez M., Gonzalez J. R., Algora C. and Espinet P. (2010). ‘Novel accelerated
testing method for III–V concentrator solar cells’, Microelectronics Reliability 50,
1880–1883.
O’Neill M., McDanal A., Walters R. and Perry J. (1991), ‘Recent development in lin-
ear Fresnel lens concentrator technology including the 300 kW 3M/Austin Sys-
tem and the 20 kW PVUSA system and the concentrator initiative’, Conf. Record
22nd. IEEE Photovoltaic Specialists Conf., Las Vegas. IEEE Press, Piscataway, NJ,
pp. 523–528.
Optics.org (2013), ‘Spectrolab beats unconcentrated PV efficiency record’, http://optics.
org/news/4/4/16, 15 April.
Randall D. E. and Grandjean N. R. (1982), ‘Correlation of insolation and wind data for
SOLMET stations’, SAND82-0094, Sandia National Laboratories.
Ross R. T. and Nozik A. J. (1982), ‘Efficiency of hot-carrier solar energy converters’, J. Appl.
Phys. 53, 3813–3818.
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch11 page 569
Rubio F., Martínez M., Hipólito A., Martín A. and Banda P. (2010a), ‘Status of CPV technol-
ogy’, Proc. 25th European Photovoltaic Solar Energy Conf., Valencia, pp. 1008–1011.
Rubio F., Martínez M., Martín A., Hipólito A. and Banda, P. (2010b), ‘Evaluation parameters
for CPV production’, 6th Int. Conf. on Concentrating Photovoltaic Systems, American
Institute of Physics, Freiburg, Germany, pp. 252–255.
Rubio F., Martínez M., Sánchez D., Aranda R. (2011), ‘Results of three years of CPV
demonstration plants in ISFOC’, Conf. Record 37th IEEE Photovoltaic Specialists
Conf., Seattle> IEEE Press, Piscataway, NJ, pp. 3543–3546.
Sala G. (1989), ‘Cooling of solar cells’, Ch. 8 in Luque A., ed. Solar Cells and Optics in
Photovoltaic Concentration, Adam Hilger, Bristol.
Sala G. and Anton I. (2011), ‘Photovoltaic Concentrators’, in Luque A. and Hegedus S. (eds),
Handbook of Photovoltaic Science and Engineering, John Wiley & Sons, Chichester,
UK.
Sala G., Araújo G. L., Luque A., Ruiz J. M., Coello A., Lorenzo E., Chenlo F., Sanz J.
and Alonso A. (1979), ‘The Ramón Areces concentration photovoltaic array’, Sun
II: Proc. ISES International Solar Energy Society Silver Jubilee Congress, Pergamon
Press, Atlanta, GA, pp. 1737–1741.
Sala G., Arboiro J. C., Luque A., Zamorano J. C., Miñano J. C. and Dramsch C. (1996),
‘The EUCLIDES prototype: an efficient parabolic trough for PV concentration’, Conf.
Record 25th IEEE Photovoltaic Specialists Conf., Washington D. C. IEEE Press, Pis-
cataway, NJ, pp. 1207–1210.
Salim A. and Eugenio N. (1990), ‘A comprehensive report on the performance of the longest
operating 350 kW concentrator photovoltaic power system’, Solar Cells 29, 1–24.
Sinton R., Kwark Y. and Swanson R. M. (1985), ‘23-percent efficient Si point contact
concentrator solar-cell’, IEEE Trans. Electron Devices 32, 2553.
Shockley W. and Queisser H. J. (1961), ‘Detailed balance limit of efficiency of p-n junction
solar cells’, J. Appl. Phys. 32, 510–519.
Solar Industry (2013), ‘CPV industry converges on standard rating conditions’, 7 March.
Terrón M. J., Davies P. A., Oliván J., Alonso J. and Luque A. (1992), ‘Deep emitter silicon
solar cells for concentrated sunlight and operation with light confining cavities’, Proc.
11th European Photovoltaic Solar Energy Conf., Montreux, Switzerland. Harwood
Academic Publishers, Chur, pp. 233–236.
Tobías I. and Luque A. (2002), ‘Ideal efficiency of monolithic, series-connected multijunc-
tion solar cells’, Progr. Photovoltaics 10, 323–329.
Vázquez, M., Algora C., Rey-Stolle I. and González J. R. (2007), ‘III–V concentrator solar
cell reliability prediction based on quantitative LED reliability data’, Progr. Photo-
voltaics 15, 477–491.
Victoria M., Dominguez C., Antón I. and Sala G. (2009), ‘Comparative analysis of different
secondary optical elements for aspheric primary lenses’, Optics Express 17, 6487–
6492.
Welford E. T. and Winston R. (1978), The Optics of Non-Imaging Concentrators, Academic
Press, New York.
Winston R. and Gordon J. R. M. (2005), ‘Planar concentrators near the étendue limit’, Optics
Lett. 30, 2617–2619.
Winston R. and Welford E. T. (1979a), ‘Geometrical vector flux and some new non-imaging
concentrators’, J. Opt. Soc. Amer. 69, 532–536.
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch11 page 570
Winston R. and Welford E. T. (1979b), ‘Ideal flux concentrators as shapes that do not
disturb the geometrical vector flux field: a new derivation of the compound parabolic
concentrator’, J. Opt. Soc. Amer., 69, 536–539.
Wojtczuk S., Chiu P., Zhang X., Derkacs D., Harris C., Pulver D. and Timmons M. (2010),
‘InGaP/GaAs/lnGaAs 41% concentrator cells using bi-facial epigrowth’, 35th IEEE
Photovoltaic Specialists Conf., IEEE, Honolulu, Hawaii, pp.1259–1264.
Yamaguchi M. and Luque A. (1999), ‘High efficiency and high concentration in photo-
voltaics’, IEEE Trans. Electron Devices 46, pp. 2139–2144.
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch12 page 571
CHAPTER 12
12.1 Introduction
The electricity from photovoltaic (PV) cells can be used for many different appli-
cations, from power supplies for small consumer products to large power stations
feeding electricity into the grid. Previous chapters in this book have discussed
different cell technologies and the optimisation of cell structures to achieve high
efficiency of conversion from light to electricity. In this chapter, we will address
the aspects that allow us to take those photovoltaic cells and incorporate them into
a system delivering a required service.
The chapter concentrates on the use of the most common types of photovoltaic
cells available commercially, described mainly in Chapters 3 to 6, and on typical
system applications, including both stand-alone and grid-connected options. The
stand-alone system can be defined as one where the PV array, together with other
system components, is designed to be the only power source for the load. By
contrast, the grid-connected PV system operates in parallel with the normal grid
distribution system and feeds any electricity not required for local loads into the
grid. This chapter considers the design and use of flat-plate PV systems. Readers
are referred to Chapter 11 for a discussion of the issues involved in the design of
PV systems incorporating concentration of sunlight.
In the next section, we discuss the construction and performance of photo-
voltaic modules. Individual solar cells must be connected to provide an appropriate
electrical output and then encapsulated so as to protect the cells from environmen-
tal damage, particularly from moisture. The design of the module depends on the
application for which it is to be used and an expansion of those applications in
571
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch12 page 572
572 N. M. Pearsall
recent years has led to a range of alternative module designs, including the use
of variable transparency and different electrical configurations. Finally, in Sec-
tion 12.2, we discuss module testing, including the establishment of rated output
and long-term performance.
In Section 12.3, we turn to the design of PV arrays, including electrical config-
uration, optimum tilt angle and orientation, protection from shading and mounting
aspects. The variation in performance expected from different array configurations
will be discussed.
The next Section 12.4 discusses the PV system, commencing with the rest
of the system components, usually referred to as the balance-of-system (BOS)
equipment. The BOS portion of the system differs substantially according to the
application and the use of the electricity produced by the PV array. In this section
we discuss the requirements of equipment to be included in a PV system, testing
and standardisation, issues of power conditioning and sizing of the PV system to
meet the required application. Finally, we will discuss applications of PV systems
including current implementation, some aspects of integration into the electrical
network and some environmental considerations.
Glass
Encapsulant
Tedlar
(a)
Superstrate
(b)
Figure 12.1 Schematic of module construction for: a) crystalline silicon cells — exploded
view showing the different layers which make up the module; b) thin-film cells.
574 N. M. Pearsall
Current = I cell
Voltage = 3 x Vcell
V oc Voltage 3 x V oc
0
Current
Single cell
3 cells in series
I sc
all the top contacts are connected together, as are all the bottom contacts. In both
cases, this results in just two electrical connection points for the group of cells.
Series connection
Figure 12.2 shows the series connection of three individual cells as an example;
the resultant group of connected cells is commonly referred to as a series string.
The current output of the string is equivalent to the current of a single cell and
the voltage output is the sum of the voltages from all the cells in the string (in this
example, since there are three cells connected, the voltage output is equal to 3Vcell ).
It is important to have well-matched cells in the series string, particularly with
respect to current. If one cell produces a significantly lower current than the other
cells (under the same illumination and temperature conditions), then the string will
operate at that lower current level and the remaining cells will not be operating at
their maximum power points. This could also happen in the case of partial shading
of a string (further discussed in Section 12.3.5).
Parallel connection
Figure 12.3 shows the parallel connection of three individual cells as an example.
In this case, the current from the cell group is equivalent to the sum of the
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch12 page 575
Current = 3 xI cell
Voltage = Vcell
Voltage V oc
0
I sc
Current
Single cell
3 cells in parallel
3 x I sc
current from each cell (here, 3Icell ), but the voltage remains equivalent to that of a
single cell.
As before, it is important to have the cells well matched in order to gain
maximum output, but this time the voltage is the important parameter since all
cells must be at the same operating voltage. If the voltage at the maximum power
point is substantially different for one of the cells, then this will force all the cells
to operate off their maximum power point and the output power will be reduced
below the optimum.
576 N. M. Pearsall
was most readily achieved by the series connection of about 36 crystalline silicon
cells. This produces a rated voltage at maximum power point voltage of around
18 V. Although the voltage reduces with reducing light intensity and increasing
temperature, this configuration allows the module voltage to remain over 12 V,
and therefore capable of charging a 12 V battery, for a wide range of operating
conditions.
Many more PV modules are now used in grid-connected applications, where
the output voltage is not constrained by battery requirements. However, most crys-
talline silicon power modules use between 36 and 72 series-connected cells since
this provides a useful voltage and reasonable module size, which can be handled
and transported. In some cases, a 72-cell module will be configured as two strings
of 36 cells each, either then connected in parallel or with separate terminals so
as to allow flexibility in module connection. In general, larger modules are used
for large ground-mounted systems and architectural applications, whilst smaller
modules are used for small systems to allow greater flexibility in electrical design.
The physical arrangement of cells in the module is generally designed to min-
imise the area, consistent with the minimum spacing between the cells required to
prevent electrical interactions and at the edges of the module to provide environ-
mental protection. Therefore, most modules are rectangular in shape and a typical
configuration is four rows of nine cells per row to obtain a 36-cell module. Typi-
cal module sizes are 0.5–2 m2 in area and the module should also be suitable for
transportation and generally light enough to be lifted by one or two people for ease
of installation.
In the case of thin-film modules, the same design principle is adopted in terms
of voltage enhancement by series connection, although the number of cells in the
string will differ since the voltage per cell is dependent on the cell technology. This
is accomplished by the electrical connection of the cells during fabrication using
a series of patterning steps. As new materials are developed, the module structure
will be adapted for the characteristics of that material (see, for example, Chapter 9
for a discussion of dye-sensitised solar cells). Some modifications of the structure
and manufacturing process can also be expected for developments of existing cell
types (e.g. rear-contact silicon cells).
4.5
60
4
3.5 50
3
40
Current (A)
Power (W)
2.5
2 30
1.5 20
Current–voltage curve
1 Power curve
10
0.5
0 0
0 5 10 15 20 25
Voltage (V)
Figure 12.4 I –V characteristic of a crystalline silicon module, with the variation of power
with voltage also shown. This illustrates the position of the maximum power point. In this
case, the module has a peak power output of around 65 W.
series and shunt resistances refer to the whole module and are dependent on the
type, number and electrical connection method of the cells.
Figure 12.4 shows an I –V characteristic together with the power curve, to
illustrate the position of the maximum power point. Owing to mismatch between
the characteristics of the component cells and to an increased overall series resis-
tance, the module will typically have a reduced fill factor as compared with its
constituent cells.
578 N. M. Pearsall
on the operating conditions and whether differences in cell performance are light or
temperature induced. Where possible — for example, for crystalline silicon cells —
manufacturers usually sort their cells into batches according to their performance
and use cells from the same batch to construct the modules. In this way, mismatch
losses are minimised.
The module efficiency is related to the total area of the module in the same
way that the efficiency of a cell is related to the total area of the cell. Because
it is necessary to have the cells physically separated, the module area is always
larger than the sum of the cell areas and therefore the module efficiency is always
lower than the cell efficiency. The resulting reduction in efficiency depends on
the configuration of the module and is defined by the packing density (ratio of
cell area to module area). Typically, a fully packed crystalline silicon module will
have a packing density of around 85% and so, if it uses 17% efficient cells, the
module efficiency will be around 14.5%. While most modules have maximum
packing density in order to produce the highest output possible per unit area, some
architectural applications require the modules to have a higher level of transparency,
for example to provide a combination of electrical generation, shading and daylight
access. In this case, the cells are more widely separated and the resulting module
efficiency is reduced accordingly.
For thin-film cells, the reduction in efficiency is lower because the strip cells
are separated only by the contact strip. In this case, the mismatch between cell
performances can be more important since it is not possible to sort and select the
cells as for the crystalline devices. Since the mismatch arises from variations in
the production process across the surface of the module, it is important to control
the uniformity of all processes.
The performance of the module is also a function of its operating temperature
and hence the rated efficiency is quoted at a standard temperature of 25◦C. The mod-
ule voltage reduces with increasing temperature and, although the current increases
slightly, the overall effect is that the efficiency reduces as the temperature rises. The
amount of the change depends on the cell type and structure, with crystalline silicon
cells typically losing about 0.4–0.5% of their output per degree Celsius rise in tem-
perature. Higher bandgap cells have a lower temperature coefficient; for example,
thin-film silicon reduces by only about 0.2% per degree Celsius. However, thin-film
silicon modules also exhibit a thermal dependence due to annealing of the light-
induced (Staebler–Wronski) degradation and this acts in the opposite direction.
So, their overall temperature coefficient can be zero or even slightly positive over
some temperature ranges. This depends on cell structure and operating conditions.
The operating temperature varies as a function of the climatic conditions of
ambient temperature and incident sunlight and also depends on the module design
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch12 page 579
and mounting. Both these latter factors affect the ability of the module to lose heat
and hence determine the operating temperature under given climatic conditions.
A measure of the effect of module design is given by the nominal operating cell
temperature (NOCT) of the module. This is the operating temperature of the cell
junction under normal incidence sunlight at an intensity of 800 W m−2 , when the
module is held at open circuit and mounted on an open rack, the ambient temper-
ature is 20◦ C and the wind speed is 1 m s−1 (IEC, 2007).
Figure 12.5 Typical polycrystalline silicon module (photograph courtesy of Sharp Elec-
tronics Europe GmbH). The square cells allow a high packing density.
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch12 page 580
580 N. M. Pearsall
In modern crystalline silicon modules, the front surface is usually glass, tough-
ened to provide physical strength and with a low iron content to allow transmission
of short wavelengths in the solar spectrum. The rear of the module can be made from
a number of materials. One of the most common is Tedlar (see Fig. 12.1), although
other plastic materials can also be used. If a level of transparency is required, then
it is possible to use either a translucent Tedlar sheet or more commonly a second
sheet of glass. The glass-glass structure is popular for architectural applications,
especially for incorporation into glazed façades or roofs.
The glass-Tedlar module is usually fabricated by a lamination technique. The
electrically connected cells are sandwiched between two sheets of encapsulant, for
example EVA (ethylene vinyl acetate), and positioned on the glass sheet that will
form the front surface of the module. The rear plastic sheet is then added and the
whole structure is placed in the laminator. Air is removed and then reintroduced
above a flexible sealing membrane to provide pressure on the module to expel air
from the structure. The module is heated and the encapsulant flows and surrounds
the cells. Additional encapsulant material is included at the module perimeter to
ensure complete sealing of the module edges.
In the thin-film module, the glass substrate on which the cell is deposited
is often used as the front surface of the finished module. The module is then
laminated in the same way as for crystalline modules although only a single layer
of encapsulant is required. Lower temperatures are often used to avoid damage to
the cells. If the substrate used for the cell fabrication is positioned at the bottom
of the module, a second glass layer is used to give the transparent front surface
required. In some cases, the thin-film solar cells are deposited on a flexible substrate
(metal foil or plastic). In all cases, particular care must be taken with edge sealing
since all thin-film cells tend to be badly affected by the ingress of moisture. During
manufacture, a clear gap must be left around the edge of the deposited cell area for
proper sealing of the module.
The electrical connections to the module are made via a junction box, usually
fixed to the rear of the module, or by flying leads. These typically exit the module
through the rear Tedlar sheet. In glass-glass modules, the leads may exit through one
edge of the module to avoid drilling holes in the glass sheet. The points at which the
electrical connections are brought out of the module are sealed to prevent moisture
ingress.
The module will exhibit the highest efficiency when the maximum amount of
the light falling on the module is incident on the cells. Light which is incident on the
spaces between cells or at the module edge is either reflected or converted to heat.
Most power modules use the minimum cell spacing, which is 2–3 mm between the
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch12 page 581
cell edges. This gap is to prevent any problems with electrical shorting between
cells.
The shape of monocrystalline silicon cells is usually described as pseudo-
square, where the cell is cut from a circular wafer and is square apart from the
cut-off corners. Polycrystalline silicon cells are often truly square, depending on
the manufacturing technique of the material. Thin-film cells are deposited in strips,
usually of around 1 cm in width and running the length of the module, although
dimensions can vary depending on cell properties.
In good sunlight conditions and with an ambient temperature of 25–30 ◦ C,
the module usually operates at temperatures in the region of 50–80 ◦ C; the actual
value depends on the module mounting. While these operating temperatures are
not excessive, the difference in thermal expansion of the various components must
be taken into account. Also, allowance must be made for the higher temperatures
experienced during manufacture, albeit for a much shorter time. The cell stringing
allows for some differential expansion in the length of ribbon between each cell.
The electrical connection is also made in two or three places on each cell to allow
for any problems with thermal expansion and other stresses during manufacture or
operation.
The ideal module would also provide good heat transfer in order to keep the cell
temperature as low as possible. However, the encapsulant is required to provide
electrical isolation and physical protection, so a high heat transfer coefficient is
not always possible. The operating temperature is also influenced by the module
encapsulation: glass-glass structures usually run at a higher temperature than glass-
Tedlar modules under similar conditions. The colour of the rear Tedlar film also has
some influence. For example, a module with a white Tedlar backing will reject more
heat than one with a black Tedlar backing, allowing it to operate at higher efficiency.
The module is often provided with a metal frame in order to make it straightfor-
ward to fix to a support structure, although this is less usual for building integrated
applications, where use can be made of traditional mounting methods for glazing.
The frame also provides some protection of the module edges in operation.
582 N. M. Pearsall
the module in operation. Testing conditions have been defined for accelerated life
testing. These include thermal cycling, hail impact, humidity-freeze and electrical
isolation tests and are detailed in IEC standard 61215 for crystalline silicon modules
(IEC, 2005) and standard 61646 for thin-film modules (IEC, 2008). Whether a
module meets the standard is determined by setting maximum limits for change in
output and visual faults after each test. As the applications for photovoltaic systems
change, the operating conditions and lifetime requirements also change. Therefore,
the performance standards are regularly reviewed, updated and added to in order
to meet the requirements of the evolving industry.
However, it is now common for bypass diodes to be incorporated into the module
structure at the manufacturing stage. Several diodes may be used, each protecting
different sections of the module, with the diode typically being connected across
twelve to eighteen cells. This integration reduces the need for extra wiring, although
it makes it more difficult to replace the diode if it fails. Where the likelihood of
shading is very low and the system voltage is also low, modules without bypass
diodes are sometimes used.
In systems where shading may reduce the output of one of the strings sub-
stantially below that of the others, it can also be advantageous to include a
blocking diode connected in series with each string. This prevents the current
from the remainder of the array being fed through the shaded string and causing
damage.
584 N. M. Pearsall
as those in northern Europe, the maximum output is usually obtained for tilt angles
of the latitude angle minus about 10–15◦ , due to the effect of both day length
and sunlight levels in the summer. The optimum tilt angle is also affected by the
proportion of diffuse radiation in the sunlight, since diffuse light is only weakly
directional. For locations with a high proportion of diffuse sunlight, the effect of
tilt angle is therefore reduced.
However, although this condition will give the maximum output over the year,
there can be considerable variation in output with season. This is particularly true in
high-latitude locations, where the day length varies significantly between summer
and winter. Therefore, if a constant or reasonably constant load is to be met or,
particularly, if the winter load is higher than the summer load, then the best tilt
angle may be higher in order to boost winter output.
Prevailing weather conditions can also influence the optimisation of the array
orientation if they affect the sunlight levels available at certain times of the day.
Alternatively, the load to be met may also vary during the day and the array can
be designed to match the output with this variable demand by varying the azimuth
angle. For example, for air-conditioning loads peaking in the afternoon, it may
be advantageous to orient the array to the west of south to boost the afternoon
output.
Notwithstanding the ability to tailor the output profile by altering the tilt and
azimuth angles, the overall array performance does not vary substantially for small
differences in array orientation. Figure 12.6 shows the percentage variation in
annual insolation levels for the location of London as tilt angle is varied between 0
and 90 degrees and azimuth angle is varied between −45◦ (south east) and +45◦
(south west). The maximum insolation level is obtained for a south-facing surface
at a tilt angle of about 35◦ , as would be expected for a latitude angle of about 51◦ N.
However, the insolation level varies by less than 10% with changing azimuth angle
at this tilt angle. A similarly low variation is observed for south-facing surfaces for
+/−30◦ from the optimum tilt angle.
The final point to consider when deciding on array orientation is incorporation
into the support structure. For building-integrated applications, system orientation
is also dictated by the nature of the roof or façade in which it is to be incorporated.
It may be necessary to trade off the additional output from the optimum orientation
against the additional costs incurred to accomplish this. Aesthetic issues must also
be considered.
100
95
Percentage of maximum solar radiation
95 –100
90 –95
90 85 –90
80 –85
75 –80
85 70 –75
65–70
80
75
45
70 30
15
0
65 Azimuth angle
–15
0 5 10
15 20 25 30 35 –30
40 45 50
55 60 65 –45
70 75 80
Tilt angle (degrees) 85 90
Figure 12.6 Percentage variation of annual global sunlight levels as a function of tilt angle
and azimuth angle. The calculations were carried out for the location of London using
Meteonorm Version 4.0.
flat-plate (non-concentrating) arrays are designed to track the path of the Sun. This
can account fully for the Sun’s movements during the day and by season by tracking
in two axes, or partially by tracking only in one axis, from east to west.
The output gain achieved by tracking for a flat-plate system depends on the
latitude, the climatic conditions (especially the ratio of direct to diffuse radiation),
the type of photovoltaic module used and the local terrain. Because the tracked
system receives more sunlight across the day, it will generally operate at a higher
temperature than the fixed array and so the efficiency will be reduced by an amount
determined by the array mounting and the temperature dependence of the type of
module being used. However, this is usually more than offset by the increased light
level received. In terms of climate, higher gains are seen where there is a high
proportion of direct sunlight.
Huld et al. (2008) have predicted the gain from two-axis tracking for sites
in Europe, compared with arrays fixed at the optimum angle for the site, in both
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch12 page 586
586 N. M. Pearsall
cases based on crystalline silicon modules. They show that, on an annual basis,
the potential gain is largest in northern latitudes (>60◦ N) and is in the region of
40–60%. This is because of the long days in the summer, when the movement of the
Sun takes it behind the fixed-plate system at the beginning and end of the day. Gains
of 30–40% are observed in southern Europe (latitudes <45◦N) due to the high direct
component in the sunlight at these locations. Lower gains of 20–30% would be
expected in central Europe owing to higher proportions of diffuse sunlight. These
values assume no restrictions due to the local terrain (e.g. mountains) preventing
tracking throughout the day.
Of course, a tracking system incurs extra costs both for the initial installation
and for maintenance, due to the required movement. Large installations will require
multiple tracking pedestals, with the number dictated by the weight of the sub-array
that can be accommodated by each pedestal. In general, a large installation of a two-
axis tracked system would also require more land area than a fixed system of the
same capacity, due to the need to avoid shading of the sub-arrays by neighbouring
ones as they move. These extra costs need to be considered against the additional
output in selecting whether to use a fixed or tracking approach for a flat-plate
system.
For concentrator systems, described in Chapter 11, the system must track the
Sun to maintain the concentrated light falling on the cell. The accuracy of tracking,
and hence the cost of the tracking system, increases as the concentration ratio
increases.
12.3.5 Shading
Shading of any part of the array will reduce its output, but this reduction will vary
in magnitude depending on the electrical configuration of the array. Clearly, the
output of any individual cell or module that is shaded will be reduced according
to the reduction of light intensity falling on it. However, if this shaded cell or
module is electrically connected to other cells and modules that are unshaded, their
performance may also be reduced since this is essentially a mismatch situation.
For example, if a single module in a series string is partially shaded, its current
output will be reduced and this will then dictate the operating current of the whole
string. If several modules are shaded, the string voltage may be reduced to the point
where the open-circuit voltage of that string is below the operating point of the rest
of the array, and then that string will not contribute to the array output. If this is
likely to occur, it is often useful to include a blocking diode for string protection,
as discussed earlier.
Thus, the reduction in output from shading of an array can be significantly
greater than the reduction in illuminated area, since it results from the loss of
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch12 page 587
output from:
588 N. M. Pearsall
DC
loads
PV array
Charge Battery
controller bank
= AC
loads
~
Inverter
DC
loads
Charge
controller
PV array
Battery
~ bank = AC
~ loads
=
Motor Rectifier Inverter
generator
systems are then known as ‘hybrid’ systems. Hybrid systems can be used in both
stand-alone and grid-connected applications but are more common in the former
because, provided the power supplies have been chosen to be complementary, they
allow reduction of the storage requirement without increased probability of not
being able to meet the load, as discussed in Section 12.4.6. Figures 12.7 to 12.9
show schematic diagrams of the three main system types.
DC
loads
Charge
controller
PV array
Battery
~ bank = AC
~ loads
=
Motor Rectifier Inverter
generator
Figure 12.9 Schematic diagram of hybrid system incorporating a photovoltaic array and a
motor generator (e.g. diesel or wind).
(if required) and load equipment. It is particularly important to include the load
equipment for a stand-alone system because the system design and sizing must take
the load into consideration. By convention, the array components are split into the
photovoltaic part (the PV modules themselves) and the balance-of-system (BOS)
components. In the remainder of this section, we briefly discuss the most common
system components and their role in the system operation, with some examples
of typical performance. There are many different options for BOS equipment,
depending on the detail of the system, and it is only possible to give a general
overview here.
Power conditioning
Most systems incorporate electrical conditioning equipment to ensure that the
system operates under optimum conditions. In the case of the array, the highest
output is obtained for operation at the maximum power point. Since the voltage
and current at maximum power point vary with both solar irradiance level and
temperature, it is usual to include control equipment to follow the maximum power
point of the array, commonly known as the maximum power point tracker (MPPT).
The MPPT is an electrical circuit which can control the effective load resistance
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch12 page 590
590 N. M. Pearsall
that the PV array sees and thus control the point on the I –V characteristic at which
the system operates.
One of the most common tracking methods is the perturb and observe (P&O)
approach. At regular intervals, the power level is checked for a small change in
operating voltage on either side of the current operating point. If a gain in power is
observed in one direction, then the MPPT moves the operating point in that direction
until it reaches the maximum value. It is also possible to calculate the expected
maximum power point based on measurement of open-circuit voltage or knowledge
of the operating conditions of the module. These methods can achieve a more rapid
determination of the maximum power point but require additional measurements
to be made. For grid-connected systems, the MPPT is often incorporated into the
inverter for ease of operation, although it is possible to obtain the MPPT as an
independent unit.
Although many stand-alone systems use an MPPT approach, some systems
operate at constant voltage, so as to allow battery charging or direct current (DC)
loads to be met at a specific voltage level. It may be necessary to include a DC–DC
converter to change the voltage level of the output of the array to that required
for input to the load. For systems with storage batteries, it is also usual to include
charge control circuitry, in order to control the rate of charge and so prevent damage
to the batteries and extend their lifetime.
Inverter
If the PV system is to supply alternating current (AC) loads or feed power directly
into the AC distribution grid, then an inverter must be included to convert the DC
output of the PV array to the AC output required. As with PV systems, inverters
can be broadly divided into two types, these being stand-alone and grid-connected
(sometimes referred to as line-tied).
The stand-alone inverter is capable of operating independently from the util-
ity grid and uses an internal frequency generator to obtain the correct output fre-
quency (usually 50/60 Hz). By contrast, the grid-connected inverter must integrate
smoothly with the electricity supplied by the grid in terms of both voltage and
frequency and therefore uses the grid as a reference. Some modern inverters have
the capability to change from stand-alone to grid-connected mode. The output volt-
age of the inverter is chosen according to the load requirements, e.g. 220–230 V
single-phase for European domestic appliances. However, if the electricity from
the PV system is to be fed directly into the supply of a large office building, for
example, a 415 V three-phase output may be chosen. For large PV installations
feeding directly into the distribution grid, higher voltages may also be chosen. The
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch12 page 591
output voltage of the inverter is selected to match the grid voltage in the location
at the network level required.
The input voltage to the inverter depends on the design of the PV array, the
output characteristics required and the inverter type. Stand-alone systems com-
monly operate at 12, 24 or 48 V, since the system voltage is determined by the
storage system, whereas grid-connected inverters usually operate at significantly
higher voltages (between 150 and 1000 V, depending on the system size).
The shape of the output waveform is important because some loads can over-
heat or be damaged if a square wave output is used. True sine wave or quasi-sine
wave (or modified sine wave) outputs are generally more costly but are much
more widely applicable. Most modern stand-alone inverters provide a modified
sine wave output, whilst grid-connected inverters should have a sine wave output
with a very low harmonic content. Limits on the harmonic distortion are defined
in the regulations for grid connection.
While most systems use an inverter connected to a group of modules, it is
also possible to obtain modules with an integrated inverter, positioned on the rear
of a module and converting the electrical output from that single module. This
module–inverter combination is sometimes referred to as an AC module. There are
some advantages in system design, such as the use of AC wiring for most of the
power transmission and reduced losses for non-uniform systems (e.g. where there is
shading) since each module has individual maximum power point tracking. Indeed,
the system designer now has a wide choice of inverter configurations, from a single
inverter per module via inverters connected to each string to a single central inverter
per system. A recent development has been the use of individual maximum power
point trackers, attached to one or two modules, with the output from a collection of
modules then being sent to a single inverter for DC/AC conversion. The choice of
system configuration depends on the uniformity of irradiance (particularly whether
there are shading problems), the overall efficiency of the different options and the
overall cost.
Inverters for PV systems are designed to have high conversion efficiency and,
although this varies with inverter design, values in the range of 95–97% for maxi-
mum efficiency are now common. The efficiency varies with the operating point of
the inverter, but usually reaches its maximum between 30 and 50% of rated capac-
ity and shows only a small decrease as the power level increases. However, the
efficiency generally reduces substantially at power levels below about 10% of full
power. Because of the variation of efficiency with operating conditions, it is also
usual for European inverter manufacturers to quote the Euro-efficiency value of
their unit. This is calculated by considering the efficiency at several input levels and
producing a weighted average according to a defined set of operating conditions.
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch12 page 592
592 N. M. Pearsall
Recently, it has been argued that this method should be refined to also take account
of the maximum power point voltage range over which the inverter operates
(Bletterie, 2011), although this modification has not yet been widely adopted.
In locations in the middle and north of Europe, the performance at low light
levels (and hence low power levels) can have a significant effect on the overall
system efficiency. Thus, it is usual to size the inverter at about 80% of the array
capacity so that high inverter efficiencies are maintained at lower power levels.
This means that the very high power levels are sacrificed since they are out of the
range of operation of the inverter, but the balance of low and high power operation
is usually such that it is more advantageous to use a reduced inverter size. This may
not be the case for systems that experience a significant proportion of high power
levels due to cold, clear weather conditions. Although all locations experience low
light level conditions, at least for fixed flat plate systems, at the end of the day,
southern European locations operate more often under high irradiance conditions
and therefore it may be most beneficial to derate the inverter by 5–10% only. The
nature of the installation should also be taken into account. For example, the solar
irradiation received on an array mounted on a vertical façade will be less than
that for the same array mounted at the optimum tilt angle in the same location.
Therefore a smaller inverter would be advisable for the façade system.
When the inverter is grid-connected, it is essential to ensure that the system
will not feed electricity back into the grid when there is a sustained fault on the
grid distribution system. This problem is known as islanding, and safeguards are
required in order to provide protection for equipment and personnel involved in the
correction of the fault. Islanding is usually prevented by closing down the inverter
when the supply from the grid is outside certain limits. The allowable limits vary
from country to country but are usually around +/−2% in voltage and frequency.
Requirements for prevention of islanding for systems are detailed in the connection
regulations for each country.
In recent years, there has been a growing amount of embedded generation (i.e.
electricity generation systems sited towards the consumer end of the distribution
system) connected to the grid in some countries. This includes wind, photovoltaic
and other systems. Problems can occur if all this generation turns off when there is
a transient fault on the grid and in some countries grid operators are now requiring
a fault ride-through capability on the inverter to address this issue. This means that
the embedded generation will continue to operate for a short period to maintain
continuity of supply. If the grid fault persists beyond a specified time, the inverter
will then close down until the fault is cleared.
Although not yet fully realised, the inverter also has the capability to contribute
to the safety and stability of the grid by a variety of means. Most inverters currently
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch12 page 593
operate at unity power factor, but there is the possibility to supply reactive power
when required by the grid, at the penalty of some loss of active power output.
The inverter also often provides system monitoring and could be incorporated
into a smart grid configuration, supplying information on the PV system status
and output. Some inverters also supply integral storage to allow the exchange of
electricity with the grid to be managed.
Storage
For many PV system applications, particularly stand-alone, electrical power is
required from the system during hours of darkness or periods of poor weather.
In this case, storage must be added to the system. Typically, this is in the form of
a battery bank of an appropriate size to meet the demand when the PV array is
unable to provide sufficient power.
The storage capacity must be chosen so as to provide sufficient security of
supply for the load and this will vary depending on how critical it is to maintain
the operation of the load. Of course, increasing storage capacity also increases the
system cost and so an appropriate balance between expenditure and performance
must be achieved.
There is a range of battery types and designs that can be used, although the most
common type is a low maintenance lead-acid battery, designed for medium depth
discharge. The design of lead-acid battery commonly used in vehicles is unsuitable
for PV applications since this only allows shallow discharge if the lifetime of the
battery is to be maintained. As previously mentioned, a charge controller should
be used so as to control the rate of charging and discharging of the battery and
to prevent complete discharge, which can seriously damage the battery. Attention
must also be paid to the location of the battery, since extremes of temperature will
also reduce operating efficiency and lifetime.
Generally, grid-connected systems do not include storage but this would be
one method to offset the variability of output from the PV system. As mentioned
in the previous section, some inverter manufacturers are beginning to offer integral
storage as an option. This adds cost but may be advantageous if the value of
deferring use of the electricity from the PV system is higher than that of exporting
it to the network. It is likely that storage will become a more common feature of
grid-connected systems as the level of penetration on the network increases.
Load equipment
The nature of the load equipment will determine the need for, and suitability of,
the power-conditioning equipment and the capacity of both the PV system and
the storage. The first consideration is whether the load or loads use DC or AC
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch12 page 594
594 N. M. Pearsall
electricity. In the former case, the loads can be operated directly from the PV
system or battery storage whereas AC loads will require an inverter to be included
in the system.
Where the system is grid-connected, loads are almost always AC but for
autonomous systems, a choice can be made. This choice will depend on the avail-
ability, cost and performance of the DC and AC versions of the load equipment. For
example, it is possible to obtain high-efficiency DC fluorescent lighting which, by
virtue of its superior performance compared with AC lighting, results in a smaller
capacity requirement for the PV system and hence, usually, reduced costs. In the
case of water pumping, the choice between DC and AC pumps depends on the
nature of the water supply (e.g. deep borehole or surface pump).
The requirements of the load in terms of voltage and current input range will
influence the type of power conditioning included in the system and the load profile
will determine the relative sizes of the PV system and the storage, if used. System
sizing in accordance with load details is discussed in more detail later in the chapter.
availability of the system will be low and the customer will be dissatisfied with the
equipment. Again, the cost-effectiveness is reduced.
Although many of the same principles are included in the sizing process, the
approach differs somewhat for stand-alone and grid-connected systems. Consid-
ering stand-alone systems first, the initial step is to gather the relevant information
on the location and purpose of the system. Location information includes:
596 N. M. Pearsall
So far, we have considered only average values for load and sunlight levels.
The daily sunlight levels can vary substantially and the battery storage must also
allow for providing power in periods of unusually poor weather conditions. The
length of the period to be allowed for is determined by consideration of local
weather conditions (i.e. the probability of several days of poor weather) and the
importance of maintaining power to the load. Clearly, if the system is used for
medical purposes or communications, loss of power could have serious conse-
quences, whereas for other situations, such as powering domestic TV or lighting,
it is merely an inconvenience. Since an increase in the period for which supplies
can be maintained involves an increase in the size of the PV array and/or battery
bank and hence an increase in system cost, this aspect is an important part of the
sizing exercise. Supply companies refer to this by many different terms, including
reliability, availability and loss-of-load probability (LOLP).
Clearly, the sizes of the PV array and battery bank are linked, and an increase
in the size of one can often allow a decrease in the size of the other. The sizing oper-
ation is usually an iteration of the problem to find the most cost-effective solution,
taking into account the requirements and preferences of the user. A detailed sizing
can be performed using system-sizing software, commercially available from a
variety of sources, using detailed solar data for the location and allowing a more
robust solution than using daily or monthly averages.
For a grid-connected system, it is not usually necessary to meet a particular
load but only to contribute to the general electricity supply. Some systems are
designed to feed all their output into the electricity grid whilst others (e.g. most
building-integrated systems) are designed to meet some of the load in a local area
with the rest of the requirement being supplied by the grid. These latter systems
only feed power into the grid when their output exceeds the demand of the load. The
system sizing is therefore not often governed by the size of the load, but by other
constraints such as the area available for the system and the budget available for its
purchase and installation. Therefore, most sizing packages are used to determine
potential output and to compare different options of system location and design,
rather than optimising system size as such. Modern sizing software includes options
for different mounting configurations and for the inclusion of an allowance for
shading based on details of the array surroundings.
The accuracy of the output of any simulation will depend on the accuracy of
the input data, as with all such systems. However, since there is a natural variation
in solar irradiation levels depending on climatic conditions, this must also be taken
into account in the use of results from a simulation. If average insolation data are
used, as is most common, then an average output will be obtained as a result. This
is strictly speaking only the average value over the period represented by the input
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch12 page 597
data rather than a prediction of what any future values will be. Thus it is possible
to obtain practical results from a system, which are significantly different from
the simulation results of the design process, simply because of normal climatic
fluctuations.
598 N. M. Pearsall
where all parameters are values for the same period, the system output is in kWh,
the average daily irradiance is in kWh m−2 and the array rating is in kW. The
array rating is calculated as the sum of the power ratings of the individual modules
making up the PV array. The monitoring fraction is the fraction of the period for
which monitoring data are available and have been used to determine the values of
the other parameters. The formula makes the assumption that average conditions
are experienced for the time when data are not collected and so care must be taken
with the use of PR values calculated for monitoring fractions less than 0.9.
Performance ratio values for grid-connected systems give a useful measure
of system design and operating quality with a value of between 0.8 and 0.9 being
the aim. This implies high inverter efficiency, minimum electrical losses between
the modules and the inverter, good thermal transfer to keep module temperature
to a reasonable level, and good system reliability. Performance ratio values cannot
normally be used to indicate the quality of a stand-alone system since the array may
not always be operating with maximum power point tracking and the requirement
to supply the load fully under worst-case conditions means that some potential
output is wasted under more favourable conditions. This can lead to a low PR
value even for a well-designed system. A different approach must be used for the
stand-alone system with availability of load (i.e. the provision of sufficient power
when it is needed to operate the load) being the most important aspect.
Stand-alone PV systems
Most stand-alone PV systems are used to power loads in locations that are remote
from the electricity grid supply, although they are also used in urban locations
where it may be inconvenient or expensive to use the grid supply (e.g. for low-
voltage lighting on bus shelters). Stand-alone systems generally range in size from
a single module powering a solar home system (SHS) to a few kilowatts of PV
supplying a local area grid network, although there are options for much larger
stand-alone systems for applications such as rural electrification.
The systems are autonomous and so must include some type of energy storage
to supply power in the absence of sunlight. Systems are usually categorised as
critical, meaning that there are issues relating to health or safety, or non-critical,
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch12 page 600
600 N. M. Pearsall
meaning that loss of load causes no more than inconvenience. In some cases,
systems powering applications with a high economic value (e.g. communications)
are also treated as critical applications from the design point of view. In general, a
critical system will have a higher storage capacity to allow for continued operation
through prolonged periods of low insolation compared with the average for that
location. The economics of storage dictate that, for larger systems and for those
where high reliability is paramount, a hybrid system is often used, allowing some
of the energy storage to be in the form of fuel for an internal combustion engine.
In locations where the seasonal availability of wind energy is complementary to
that of the solar irradiance, it can also be cost-effective to include a wind turbine
in the hybrid system. Other options would be biomass or small hydro-electricity
systems, depending on availability at the location.
In a small, non-critical system, such as an SHS, one or two PV modules
charge a battery during the day, and the power is used at night for a few high-
efficiency lights and a radio or small TV. A charge controller ensures that the bat-
tery is not overcharged or deep-discharged, to provide as long a battery lifetime as
possible. Battery capacity and lifetime can be substantially reduced by poor charg-
ing regimes, especially consistent discharging below the design level of charge. To
keep costs as low as possible, standard systems are sold to all users, though there
may a choice of the number of modules based on local needs and solar conditions.
System reliability largely depends on users modifying their usage of the loads to
ensure an appropriate battery charging regime, often using indicator lights on the
charge controller.
For systems that are part of safety-critical networks, for example those in
aircraft navigation aids or telecommunication systems, a very low LOLP must
be guaranteed, perhaps as little as one or two hours per year on average. The
stochastic variability of solar irradiance means that a large PV array feeding into a
large battery storage capacity must be provided to ensure that almost all periods of
poor weather can be accommodated, or additional charging must be provided from
a small internal combustion engine, usually a diesel, with a fuel store sufficient
to maintain this option with fuel deliveries under normal maintenance visits only.
The PV array and battery system should be sized so that the engine is run at full
power on a sufficiently regular basis to keep it in good condition, but so that the
overall cost is minimised, taking into account the fuel and equipment costs. The
design will vary with location in terms of both climate and cost of fuel.
Some stand-alone systems provide power for a local network for a small com-
munity. Again, the use of a hybrid system is common, although in this case the
alternative power source may be another renewable energy technology such as wind
or small hydropower to supplement the PV output. The optimum combination will
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch12 page 601
take account of the required load profile and the generation profile of the various
power sources. It is possible to commence a local network with a small PV or PV
hybrid system and then add to it as consumer demand grows and finance allows.
The first commercial use of PV cells was in space, providing the electrical
power required by satellites for their operation, and this continues to be an impor-
tant application of the technology. These systems could be described as a special
case of a stand-alone system, where the space environment must be considered in
determining the system size needed to meet the load requirement. Particular atten-
tion must be paid to the PV array size and weight, because of launch constraints,
and to the performance of the solar cells after exposure to irradiation in the space
environment.
Stand-alone PV systems are used for a wide range of applications from lighting
for rural homes to telecommunication repeater stations, from providing the power
for parking meters to maintaining the vaccine cold chain by powering refrigerators
in rural clinics, from water pumping to navigation buoys. Some systems have only
the PV array as their power source, whilst others incorporate one or more other
power supplies. Whilst the uses differ greatly, the design approach is the same.
The load is carefully considered and then the system is sized so as to achieve the
required probability that it will meet the load given the climatic conditions at the
location. In this way, PV systems can make a remarkable contribution to providing
power in difficult locations and to people who have no other source of electricity.
Grid-connected PV systems
The grid-connected system, as its name implies, is a PV system that is installed
in a location where there is also grid electricity available and a direct connection
is made to allow power to be fed to or taken from the grid. We can divide grid-
connected PV systems into two main categories, which are generally described as
centralised and distributed. Centralised systems feed electricity directly into the
grid and are usually connected at medium voltage. These systems will be discussed
later in this section. We will first consider distributed systems, which generally feed
local loads first and then export any excess to the grid. These systems are usually
connected at the level of the electricity consumer (i.e. at low voltage). The most
common example of this kind of system is the one that is mounted on or integrated
into a building.
602 N. M. Pearsall
distributed nature of both sunlight and the electrical load. The potential benefits of
the BIPV system can be summarised as follows:
Figure 12.10 Example of façade integration of photovoltaics. The photograph shows the
40 kWp PV façade on the Northumberland Building at the University of Northumbria.
The PV array is integrated into the rainscreen overcladding. This system was installed in
1994 and is an early example of façade integration. Photograph courtesy of University of
Northumbria.
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch12 page 603
Figure 12.11 Integrated PV façade using copper indium gallium diselenide thin-film
modules. Photograph courtesy of Würth Solar GmbH.
604 N. M. Pearsall
The area available for the BIPV array may be constrained by building design,
shading from surrounding structures or owner preference. Thus, the system size is
often dictated by the building’s structural details rather than its electrical loads. The
visual aspect of the system is also important and this will influence the choice of
module type, location and detailed integration method. Finally, the system design
must take into account ease of installation, maintenance and operation, and com-
pliance with local building regulations, both structural and electrical.
A fully integrated BIPV array performs at least two tasks, the generation
of electricity for use in the building and the protective functions of the external
building element, but arrays can also be designed to perform additional functions.
The most common function is shading of internal spaces, by louvre systems on
the exterior of the building, by designing the cladding so as to provide shading to
the windows at high Sun positions or by the use of semitransparent PV elements
for a roof or façade. Figure 12.12 shows an example of the use of semitransparent
modules in a glazed façade, where the cells provide both visual stimulation by
variation of the arrangement pattern and shading to reduce solar gain and glare.
In some cases the heat at the rear of the modules can also be used. For all
except the highest efficiency modules, less than 20% of the light falling on the
module is converted to electricity and, whilst a few percent is reflected, the rest
is absorbed as heat. This results in a module operating temperature that can be
Figure 12.12 Example of the use of photovoltaic modules to influence indoor lighting
patterns. The Solar Office at the Doxford International Business Park in Sunderland, UK,
has a 73 kWp array formed from semitransparent PV modules. The cell spacing is varied to
create the light effects in the inner atrium. Photograph courtesy of Akeler Developments Ltd.
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch12 page 605
606 N. M. Pearsall
The wide range of possible designs and the variation in the cost of the building
material that is being replaced, and hence the offset of the PV system cost that this
provides, makes it difficult to generalise about the cost of BIPV systems. Some
systems are less costly than the material being replaced and so the electricity
produced is essentially free except for any small additional maintenance costs.
Most systems are replacing material that would be less costly than the PV array
and so there is some residual cost to assign to the electricity produced. Nevertheless,
the reducing costs of PV modules and other system equipment means that these
distributed systems are becoming economically attractive across a wide range of
locations, especially as the costs of conventionally generated electricity rise.
optimum angle for solar capture, and spaced so that there is a low level of shading
between the rows. This requires a trade-off between the available land area and
the amount of shading at low solar elevations and the spacing can be expressed by
a parameter known as the Ground Cover Ratio. This is the ratio of the total area
of PV modules to the total land area of the site and is typically around 0.5–0.6,
although it can vary depending on the latitude and the details of the site. In some
cases, the spacing, tilt angle and even azimuth angle can be varied on different
parts of the site to take account of different topographies, the aim being to generate
the highest energy yield from the available area. Figure 12.13 shows an example
of a ground-mounted centralised PV system.
While most of the large grid-connected PV systems are fixed, some flat-plate
systems employ Sun-tracking on either one or two axes. For tracked systems, the
most common configuration is to mount a group of modules on a pedestal, to
give the required freedom of movement, similar to the configuration illustrated in
Fig. 11.29 in Chapter 11 for concentrator systems. The array field is then made
up of a multiple of pedestals up to the required system capacity. Tracking both
increases the output of the modules and modifies the output profile, keeping it
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch12 page 608
608 N. M. Pearsall
high across a longer portion of the day. As a disadvantage, a system using tracking
will require a larger area of land so as to keep the shading between the modules
to an acceptable level. The support structure and tracking system also make the
installation more costly. The performance advantage for a tracked system depends
on the climatic details at the location and is more pronounced for sites with a high
fraction of clear days across the year. The choice of whether or not to use flat-plate
tracking is therefore location-dependent. Some of the large centralised systems also
use concentrator PV technology (see Chapter 11), where the climatic conditions
are favourable.
Environmental impact
Whereas the stand-alone system is designed to meet a specific load, the motivation
for the grid-connected system is ultimately to reduce consumer energy costs and
to reduce the environmental impact of electricity production. There is an environ-
mental impact from the manufacture and disposal of PV modules and the rest of
the system components, but this is substantially lower per unit of electricity across
the system lifetime than for fossil-fuel-based generation (IPCC, 2012). Further-
more, it can be expected to reduce consistently over time, due to reducing energy
inputs in manufacturing and reducing carbon content in relation to those inputs, as
more renewable energy technologies are incorporated into the grid. The compari-
son of impact between PV systems and other renewable technologies depends on
assumptions made about the system design and location, so would require a fuller
treatment than is possible here. It can be concluded that PV systems, as part of a
portfolio of renewable technologies, can contribute significantly to the reduction
of carbon emissions resulting from electricity provision.
As the use of PV systems increases, it is also necessary to address the issue of
how to deal with the system components at the end of the system life. Photovoltaic
systems are generally straightforward to decommission and the industry, especially
in Europe, has now begun to put in place schemes for the collection and recycling
of modules at the end of their life. Within Europe, PV equipment comes under
the Waste Electrical and Electronic Equipment Directive (European Union, 2012)
and this requires the suppliers to provide a route for customers to return their used
equipment. It has been demonstrated that a large proportion of the PV module can
be recycled, with materials either being reused for manufacturing new modules
(e.g. refurbished crystalline silicon cells) or in other sectors (e.g. glass).
Network development
As the cost of PV electricity reduces towards the cost of conventionally generated
electricity in several countries, one of the technical challenges to be faced is how
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch12 page 609
to deal with large amounts of PV electricity being fed into the electricity network,
whether from distributed or centralised systems. The output from the PV system is
variable, depending on weather conditions. The output profile can be modified by
introducing storage, whether local or central, and, if sufficient capacity is installed,
by curtailment of output power to match requirements. The use of the inverter to
provide both storage and reactive power has been discussed earlier in the chapter.
Clearly, the network also has to include generation capacity to meet demand when
there is no solar output.
The ability to control the supply of electricity via the grid where there is a
high penetration level of variable renewable technologies (e.g. wind, PV) requires
knowledge of likely outputs. This is leading to advances in forecasting techniques
for system output over short and medium timescales and in the communication
systems and models required to make use of those forecasts.
12.5 Conclusions
Photovoltaic cells have social and commercial value when they are used in a system
to provide a service, whether that is the provision of electricity to meet a specific
load or to contribute to the power network serving a multitude of loads. This chapter
has given a brief overview of the technical considerations that allow the cells to
provide such a service, and of the current applications of PV systems.
Photovoltaic cells may be incorporated directly into a product and add value
to that product to the extent that their use is commercially viable. In most cases,
however, the cells are contained in a PV module, interconnected to give a required
output depending on the application. The module structure also protects the cells
from damage in transportation, installation and use. This chapter has described
the construction of PV modules and their quality assurance testing, designed to
provide a product with an assured output, reliability and lifetime when operating
in varied climatic conditions. It is these developments in module performance that
have provided the basis for the expanding market for PV throughout the world.
While the current module construction is similar in concept to that developed over
30 years ago, consistent developments in manufacturing have reduced the material
and energy requirements in manufacture and reduced the electrical losses in moving
from the individual cell to the module. Modules have also become larger and more
varied to address different user requirements. As new cell types develop and new
applications are found, PV modules will surely continue to evolve. Nevertheless,
the requirements for lifetime and durability will remain.
This chapter has also discussed the range of other equipment needed in PV
systems to give optimal operation in both stand-alone and grid-connected systems.
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch12 page 610
610 N. M. Pearsall
Here, it was possible to give only a brief overview of the equipment and the design
criteria, especially in relation to the development of inverters. As the applications
of PV systems become more widespread, especially in relation to connection to
the grid, so the BOS components are also being developed to address specific
market needs such as grid support. The interested reader is referred to the specialist
journal and conference papers for the latest information, as well as to manufacturer
information on the latest commercial products.
Photovoltaic systems are clearly well suited to the provision of electrical power
at locations that are remote from the electricity grid and where reliability is a key
requirement. This is especially true where only small amounts of power are needed
to bring major improvements in quality of life, e.g. for rural medical applications.
The main issues are the correct design of the system, the education of the user
to ensure the system is operated correctly, and the provision of financial mech-
anisms to allow the purchase of the systems (which is beyond the scope of this
chapter).
The great majority of the installed capacity of PV systems around the world is
grid-connected, with a market share estimated to be >98% at the end of 2011 (IEA
Photovoltaic Power Systems Programme, 2012). The reduction in PV module price
over the last few years and the simplicity of their installation and operation has
meant that small, distributed systems have become popular. The market promotion
schemes introduced by several countries have also encouraged the installation of
centralised PV systems and capacity has increased year on year, as described in
Chapter 13.
While PV systems only provide a very small percentage of networked elec-
tricity at the moment, there are a few locations, particularly in southern Germany,
where there is a high penetration of systems. Under the correct conditions, it has
been demonstrated that the load can be met solely from a combination of wind
and solar electricity. This situation brings challenges in relation to control and
variability of supply, but also provides the stimulus for the developments in sys-
tem hardware and output forecasting that will eventually allow PV to be a major
contributor to the world’s electricity supply.
References
Bletterie, B., Bründlinger R. and Lauss G. (2011), ‘On the characterisation of PV inverters’
efficiency — introduction to the concept of achievable efficiency’, Progr. Photovoltaics
19, 423–435.
European Union (2012), Directive 2012/19/EU of the European Parliament and of the
Council on Waste Electrical and Electronic Equipment (WEEE), Official Journal of
the European Union, 24 July.
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch12 page 611
Huld T., Šúri M. and Dunlop E. D. (2008), ‘Comparison of potential solar electricity output
from fixed-inclined and two-axis tracking photovoltaic systems in Europe’, Progr.
Photovoltaics 16, 47–59.
IEA Photovoltaic Power Systems Programme (2012), Trends in Photovoltaic Applications:
Survey Report of Selected IEA Countries between 1992 and 2011, Report IEA-PVPS
T1-21:2012.
IEC (2007), Solar Photovoltaic Energy Systems — Terms, Definitions and Symbols, IEC/TS
61836:2007.
IEC (2005), Crystalline Silicon Terrestrial Photovoltaic (PV) Modules — Design Qualifi-
cation and Type Approval, IEC 61215:2005.
IEC (2008), Thin-film Terrestrial Photovoltaic (PV) Modules — Design Qualification and
Type Approval, IEC 61646: 2008.
IPCC (2012), Renewable Energy Sources and Climate Change Mitigation, Special Report of
the Intergovernmental Panel on Climate Change, Cambridge University Press, p. 732.
Lloret A., Aceves O., Andreu J., Merten J., Puigdollers J., Chantant M., Eicker U. and
Sabata L. (1997), ‘Lessons learned in the electrical system design, installation and oper-
ation of the Mataró Public Library’, Proc. 14th European Photovoltaic Solar Energy
Conf., Barcelona, Spain. H. S. Stephens & Associates, Bedford, pp. 1659–1664.
May 2, 2013 14:6 BC: 8831 - Probability and Statistical Theory PST˙ws
CHAPTER 13
Viewed in the light of the world’s growing power needs, these gadgets are toys.
But so was the first flea-power motor built by Michael Faraday over a century ago
— and it sired the whole gigantic electrical industry.
From Prospects for Solar Power, Harland Manchester, Reader’s Digest, June 1955.
13.1 Introduction
Over the last decade photovoltaics has been one of the fastest-growing industries,
with a compound annual growth rate (CAGR) of over 50%. Solar cell and module
production have increased by more than an order of magnitude during this decade
and despite the negative impacts of the on-going economic crisis since 2009, cumu-
lative installed capacity of photovoltaic (PV) systems surpassed 100 GW1 at the end
of 2012. In 2013, new installed photovoltaic capacity could for the first time surpass
new installed wind-power capacity. This is mainly driven by rapid market expan-
sion in Asia, which will be the largest market, surpassing Europe. Photovoltaics is
on the right track to become a major electricity source within the next decade.
1 All production and installation figures cited in W in this chapter are actually Wp.
2 Solar-cell production capacities mean: in the case of wafer silicon-based solar cells, only the cells; in
the case of thin-films, the complete integrated module. Only those companies that actually produce the
active circuits (solar cells) are counted; companies that purchase these circuits and make cells are not
counted.
613
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch13 page 614
614 A. Jäger-Waldau
45
Rest of World
40 United States
Malaysia
35
Japan
Annual Production [GW]
Europe
30
Taiwan
25 PR China
20
15
10
0
2005 2006 2007 2008 2009 2010 2011 2012 2013e
Figure 13.1 World PV cell/module production from 2005 to 2012, and 2013 (estimated).
Data source: Photon International (2012), RTS Corporation (2013), PV News (2013) and
the author’s analysis.
well as the fact that some companies report shipment figures, while others report
sales and others still production figures. The data presented, collected from stock
market reports of listed companies, market reports and colleagues, were compared
to various data sources and this led to an estimate of 38.5 GW (Fig. 13.1), repre-
senting a moderate increase of about 10% compared with 2012; as manufacturing
plant utilisation began to improve in the second quarter of 2013 another moderate
increase of about 10% is expected for 2013.
Total PV production has increased almost by two orders of magnitude since
2000, and the CAGR over the last decade has been above 50%. The most rapid
growth in annual production over the last five years has occurred in Asia, where
China and Taiwan together now account for more than 70% of worldwide produc-
tion.
Publicly traded companies manufacturing products along the value chain,
installing photovoltaic electricity systems or offering related services, have
attracted a growing number of private and institutional investors. In 2012, world-
wide new investments in clean energy decreased by 11% compared with 2011 but
were still very considerable at $269 billion ( 207 billion3 ), including $30.5 bil-
lion ( 23.5 billion) corporate and government research and development spending
(Bloomberg New Energy Finance, 2013a; PEW Charitable Trusts, 2013). In 2012,
clean energy markets outside the Group of 20 (G20) grew by more than 50% to
surpass $20 billion ( 15.4 billion), while investments in the G20 countries dropped
by 16% to $218 billion ( 167.7 billion). Despite the overall decline in investments,
the decrease of renewable energy system prices more than compensated this and
allowed these investments to install a record 88 GW of new clean-energy genera-
tion capacity, bringing the total to 648 GW, capable of producing more than 1500
TWh of electricity per annum. This corresponds to 64% of the electricity generated
per annum by nuclear power plants worldwide.
For the third year in a row, solar power attracted the largest amount of new
investment into renewable energies. Despite a 9% decline in solar-energy invest-
ments, solar power attracted $137.7 billion ( 105.9 billion), which is 57.7% of all
new renewable energy investments (Bloomberg New Energy Finance, 2013b).
In contrast to Europe and the Americas, where new investments in renewable
energy decreased, new investments continued to increase in Asia/Oceania, reaching
$101 billion ( 77.7 billion) in 2012. Europe took the second place with $62.1 billion
( 47.8 billion), followed by the Americas with $50.3 billion ( 38.7 billion) (PEW
Charitable Trusts, 2013). With a 20% increase, China became the largest investor
in renewable energy with $65.1 billion ( 50.1 billion) followed by the USA with
$35.6 billion ( 27.4 billion) and Germany with $22.8 billion ( 17.5 billion). The
country with the biggest change in 2012 was South Africa, which moved up to
ninth place with $5.5 billion ( 4.2 billion).
13.1.2 Prices
The existing overcapacity in the solar industry has led to continuous downward
price pressure along the value chain and resulted in a reduction of spot market
prices for polysilicon, solar wafers and cells, as well as solar modules. Photovoltaic
module prices have decreased in price by 80% since 2008, and by 20% in 2012
alone (Bloomberg New Energy Finance, 2013b). These rapid price declines are
putting all solar companies under enormous pressure and access to fresh capital is
key to survival. It is believed that this situation will continue until at least 2015,
when the global PV market should exceed 50 GW of new installations. The increase
of polysilicon spot prices and the levelling of module prices since the beginning
of 2013 indicate that some production capacity has been taken out, and that prices
may stabilise for some time until they are back on the learning-curve slope. It
should be noted that PV system hardware costs are priced more or less the same
worldwide, and the so-called soft costs, which consist mainly of financing, permit
and labour costs, and installer/system integrator profits, are the main reason for the
significant differences which are still observed.
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch13 page 616
616 A. Jäger-Waldau
The continuation of the difficult worldwide financial situation and the fact that
support schemes are changed with little notice, undermining long-term investor
confidence, make project financing more difficult as risk premiums are added. On
the other hand, the declining module and system prices have already opened new
markets that offer perspectives for further growth of the industry — at least for those
companies with the capability to expand and reduce their costs at the same pace.
Despite the problems of individual companies, business analysts are confident
that the industry fundamentals, as a whole, remain strong and that the overall
photovoltaics sector will continue to experience significant long-term growth. For
example, in its second Medium-Term Renewable Energy Market Report, published
in July 2013, the IEA increased its predicted installed capacity to 300 GW of
cumulative PV installations by 2018 (IEA, 2013a).
Market predictions for new PV system installations in 2013 have been
upgraded regularly during the year. They range between 31 GW from Solarbuzz,
33.1 GW in the Bloomberg conservative scenario, >35 GW from IHS research and
39.6 GW in the Bloomberg optimistic scenario (Bloomberg New Energy Finance,
2012; IHS Global Demand Tracker, 2013; Solarbuzz, 2013). The wide spread of
analyst numbers derives from the different methodologies, and whether the given
numbers represent constructed/installed systems or actual grid-connected systems.
For 2014, analysts expect a further increase to over 40 GW, mainly driven by grow-
ing Asian markets. Even in the optimistic forecasts, massive overcapacities in cell
and module manufacturing will continue to exist. Depending how capacities are
calculated, overcapacity estimates for 2013 range between 60 and 70 GW.
Despite a number of bankruptcies and companies idling production lines, or
even permanent closure of production lines, the number of new entrants into the
field, notably large semiconductor or energy-related companies, has overcompen-
sated this in the past. However, the rapid changes in the sector and the difficult
financing situation make a reasonable forecast of future capacity developments
very speculative. The consequence is the continuation of price pressure in an over-
supplied market. This will accelerate the consolidation of the photovoltaics industry
and spur even more mergers and acquisitions.
The existing overcapacity is a result of very ambitious investments dating
back to 2010, triggered by the more than 150% growth of the PV market in that
year. This led to a peak in capital spend on manufacturing equipment of about
$14 billion ( 10.8 billion) in 2011. Since then, equipment spending has declined
dramatically and will probably hit the bottom with around $1–2 billion ( 0.77–1.54
billion) invested in 2013 before a moderate recovery from 2014 onwards will be
possible. This development has had a serious effect on equipment manufacturers,
all of which now need a new strategy for the PV industry. Companies with no
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch13 page 617
significant business segment outside the PV supply chain were hit the hardest,
and some of them are struggling to survive the slump until the predicted recovery
kicks in.
618 A. Jäger-Waldau
• Development of new business models for the collection, sale and distribution
of photovoltaic electricity, e.g. development of bidding pools at electricity
exchanges, virtual power plants with other renewable power producers and
storage capacities.
• Adaptation of regulatory and legal procedures to ensure fair and guaranteed
access to the electricity grid and market.
The cost of PV-generated electricity has dropped to less than 0.05 /kWh, and
the main cost component is to transport the power from the module to where it is
needed. Therefore innovative and cost-effective overall electricity system solutions
for the integration of PV electricity are needed to realise the vision of PV as a major
electricity source for everybody everywhere. More effort is needed to optimise the
non-PV costs and greater public support, especially as regards regulatory measures,
is crucial.
50
45 Rest of Europe
Annual Photovoltaic Installations/GWp
Italy
40 Spain
Germany
35 Rest of World
China
30 United States
Japan
25
20
15
10
0
2000 2005 2006 2007 2008 2009 2010 2011 2012 2013e 2014e
Figure 13.2 Annual photovoltaic installations from 2000 to 2012 and estimates for 2013
and 2014. Data sources: European Photovoltaic Industry Association, 2013; National Energy
Administration, 2013; Systèmes Solaires, 2013, and author’s analysis.
200
Cumulative Photovoltaic Installations/GWp
80
60
40
20
0
2000 2005 2006 2007 2008 2009 2010 2011 2012 2013 e 2014e
Figure 13.3 Cumulative photovoltaic installations from 2000 to 2012 and estimates for
2013 and 2014. Data sources: European Photovoltaic Industry Association, 2013; National
Energy Administration, 2013; Systèmes Solaires, 2013, and author’s analysis.
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch13 page 620
620 A. Jäger-Waldau
largest portion of the total worldwide 100 GW of installed capacity at the end of
2012.
Australia
In 2012, slightly more than 1 GW of new solar photovoltaic electricity systems were
installed in Australia, bringing the cumulative installed capacity of grid-connected
PV systems to 2.45 GW. Photovoltaic electricity systems accounted for 36% of all
new electricity generation capacity installed in 2011. As in 2011, the 2012 market
was dominated by grid-connected residential systems, which accounted for more
than 90% of the total installed. The average PV system price for a grid-connected
system fell from 6 AUD/Wp (4.29 /Wp4 ) in 2010 to 3.9 AUD/Wp (3 /Wp5 )
in 2011 and 3.0 AUD/Wp (2.15 /Wp4 ) in early 2013. Due to this, the cost of
PV-generated electricity has reached, or is even below, the average residential
electricity price of 0.27 AUD/kWh (0.19 /kWh).
India
Market estimates for PV systems in India in 2012 vary between 750 and 1000 MW,
because some statistics cite the financial year and others the calendar year. Accord-
ing to the Ministry of New and Renewable Energy (MNRE), total capacity at the
end of financial year (FY) 2012/136 was 1.9 GW grid-connected and 125 MW off-
grid capacity. The Indian Jawaharlal Nehru National Solar Mission (JNNSM) was
launched in January 2010, and it was hoped that this would give impetus to the
grid-connected market. This Mission aims to make India a global leader in solar
energy and envisages an installed solar generation capacity of 20 GW by 2020,
100 GW by 2030 and 200 GW by 2050. After only a few MW had been installed
in 2010, installations in 2011 slowly picked up, but the majority of the JNNSM
projects will come online from 2015 onwards. Market expectations for 2013 vary
between 1.00 and 1.35 GW (Bridge to India, 2013; Mercom Capital Group, 2013).
Israel
Three years after the introduction of an FiT in 2008, Israel’s grid-connected PV
market saw about 60 MW of capacity newly connected in 2012. One of the main
drives behind the development of solar energy is energy security, and Israel plans
to have about 1 GW of solar PV systems installed by the end of 2014. In August
2012, about 215 MW had been built and a further 300 MW approved (Bloomberg
New Energy Finance, 2012). The FiTs depend on the system size segment and
have individual caps. Market expectations for 2013 are in the range 150–200 MW.
622 A. Jäger-Waldau
Japan
The Japanese market experienced significant growth in 2012, increasing its domes-
tic shipments to 2.47 GW. Cumulative installed capacity increased by about 1.7 GW
to reach 6.6 GW at the end of 2012 (IEA, 2013b). By the end of February 2013,
more than 5.5 GW had received approval under the new national feed-in scheme,
which started in July 2012. The market outlook for Japan has been raised and is now
between 6.9 and 9.4 GW new installed capacity by the end of 2013 (Bloomberg
New Energy Finance, 2013d).
The consequence of the accident at the Fukujima Daiichi Nuclear Power Plant
in March 2011 was the reshaping of the country’s energy strategy. For PV power,
an official target of 28 GW was set for 2020. In July 2012, a Ministry for Economy,
Trade and Industry (METI) panel published its long-awaited plan to reform the
country’s power market. This aims to increase the share of renewable power supply
from 11% in 2011 to 25–35% by 2030.
Until 2010, residential rooftop PV systems represented about 95% of the
Japanese market. In 2011, due to a change in permit policy, large ground-mounted
systems, as well as large commercial and industrial rooftop systems, increased
their market share to about 20%.
In June 2012, METI finally issued the Ministerial Ordinances for the new FiTs
for renewable energy sources and these were adjusted in March 2013. The tariff
for commercial installations (total generated power) larger than 10kWp is 36 per
kWh for 20 years and 38 per kWh7 for residential installations (surplus power)
smaller than 10 kWp for 10 years, starting from 1 April 2013 (Ikki, 2013a).
Plan targets are 17% lower carbon dioxide emissions and 16% lower energy
consumption per unit of gross domestic product (GDP) by 2015. The total invest-
ment in the power sector under the Plan is expected to reach $803 billion ( 618
billion), divided into $416 billion ( 320 billion), or 52%, for power generation, and
$386 billion ( 298billion) to construct new transmission lines and other improve-
ments to China’s electrical grid.
In August 2012, the NEA released China’s new renewable energy five-year
plan for 2011–2015 (NEA, 2012). The new goal of the NEA calls for renewable
energy to supply 11.4% of the national energy mix by 2015. To achieve this goal,
renewable power generation capacity has to be increased to 424 GW. Hydro-power
is planned to be the main source, with 290 GW including 30 GW pumped storage,
followed by wind with 100 GW, solar with 21 GW (this target was later increased
to 41 GW) and biomass with 13 GW.
The plan estimated new investments in renewable energy of CNY 1.8 trillion
( 222 billion) between 2011 and 2015. China aimed to add a total of 160 GW of new
renewable energy capacity during the period 2011–15, in the form of 61 GW hydro,
70 GW wind, 21 GW solar8 (10 GW small distributed PV, 10 GW utility scale PV
and 1 GW solar thermal power), and 7.5 GW biomass. For 2020, the targets were
set as 200 GW for wind, 50 GW for solar (27 GW small distributed PV, 20 GW
utility-scale PV and 3 GW solar thermal power) and 30 GW for biomass.
These required investment figures to be in line with a World Bank report stating
that China needs additional investment of $64 billion ( 49.2 billion) annually over
the next two decades to implement an energy-smart growth strategy (World Bank,
2010). However, according to the report, the reductions in fuel costs through energy
savings could largely pay for these additional investment costs. At a discount rate of
10%, the annual net present value (NPV) of the fuel cost savings from 2010 to 2030
would amount to $145 billion ( 111.5 billion), which is about $70 billion ( 53.8 bil-
lion) more than the annual NPV of the additional investment costs required.
On 24 February 2012, the Chinese Ministry of Industry and Information Tech-
nology published its industrial restructuring and upgrading plan (2011–2015) for
the photovoltaic industry (MIIT, 2012). In this document, the Ministry stated that
by 2015 it expects to support only backbone enterprises, which should produce a
minimum of 50000 MT polysilicon, or 5 GW of solar cell or module production.
The Plan also projects a cost reduction of electricity generated with PV systems
to 0.8 CNY/kWh (0.098 /kWh9 ) by 2015 and 0.06 CNY/kWh (0.074 /kWh) by
2020.
624 A. Jäger-Waldau
South Korea
In 2012, about 250 MW of new PV systems were installed in South Korea, bringing
the cumulative capacity to a total of 981 MW (IEA, 2013c). Since January 2012,
Korea’s Renewable Portfolio Standard (RPS) officially replaced its earlier FiTs. For
2013, the RPS set-aside quota for PV was set at 450 MW, and this should increase to
1.2 GW in 2016. The result is an annual target of 230 MW in 2013, 240 MW in 2014,
250 MW in 2015 and 260 MW in 2016. Under the RPS, income for power generated
from renewable energy sources is a combination of the wholesale system marginal
electricity price and the sale of Renewable Energy Certificates (RECs); certificate
sales in the second half of 2012 were around 40000 KRW/MWh (0.026–0.035
/MWh10 ). These RECs are multiplied by an REC multiplier, varying between
0.7 for ground-mounted free-field systems to 1.5 for building-adapted systems, to
reflect the different costs of the different systems.
The new RPS programme obliges power companies with at least 500 MW of
generating capacity to increase their renewable energy mix from not less than 2%
in 2012 to 10% by 2022. The renewable energy mix in the Korean RPS is defined
as the proportion of renewable electricity generation to the total non-renewable
electricity generation.
Taiwan
In June 2009, the Taiwan Legislative Yuan gave its final approval to a Renewable
Energy Development Act to bolster the development of Taiwan’s green energy
industry. The goal was to increase Taiwan’s renewable energy generation capacity
by 6.5 GW to a total of 10 GW within 20 years. Promotion of all types of renewable
energy was expanded: total installed capacity of 9952 MW (accounting for 14.8% of
total power generation installed capacity), with new installed capacity of 6600 MW,
has been planned by 2025 so that the goal set by the Act can be achieved five years
early. By 2030, the target will be further increased to 12.5 GW, accounting for 16.1%
of total power generation installed capacity and capable of generating 35.6 billion
kWh of electricity. This is equivalent to the annual electricity consumption of 8.9
million households (accounting for 78% of the number of households consuming
electricity nationwide).
Between 2009 and 2012, a total capacity of 194 MW was installed, bringing
the total installed capacity to 222 MW at the end of 2012 (Ministry of Economic
Affairs, 2013). The FiTs in the first half of 2013 for rooftop systems were 8.4
NT$/kWh (0.215 /kWh11 ) for systems up to 10 kW, 7.54 NT$/kWh (0.193 /kWh)
for systems between 10 and 100 kW, 7.12 NT$/kWh (0.183 /kWh) for systems
between 100 and 500 kW and 6.33 NT$/kWh (0.162 /kWh) for systems larger than
500 kW. Ground-mounted systems had a tariff of 5.98 NT$/kWh (0.153 /kWh).
For the second half of 2013, a tariff reduction of 2.5% to 6.1% was foreseen to
reflect declining system prices.
The installation targets for 2013 were increased twice and are currently at
175 MW. This is in line with Taiwan’s new Million Solar Rooftop Programme,
which aims to achieve installed capacity of 610 MW by 2015 and 3.1 GW by
2030. However, the increased installation target for 2013 represents only 3.5% of
Taiwan’s total 2012 PV production volume, reflecting the country’s strength as a
PV exporter.
Thailand
Thailand enacted a 15-year Renewable Energy Development Plan (REDP) in early
2009, setting the target of increasing the renewable energy share of the final energy
consumption of the country to 20% by 2022. Besides a range of tax incentives,
PV systems are eligible for a feed-in premium or ‘adder’ for a period of ten years.
The original 8 THB/kWh12 (0.182 /kWh) ‘adder’ facilities in the three south-
ern provinces, and those replacing diesel systems, are eligible for an additional
1.5 THB/kWh (0.034 /kWh. This was reduced to 6.5 THB/kWh (0.148 /kWh)
for those projects not approved before 28 June 2010. The original cap of 500 MW
was increased to 2 GW at the beginning of 2012, due to the high oversubscription
of the original target. In addition to the ‘adder’ programme, projects are now being
developed with power purchase agreements (PPAs).
At the end of 2012, grid-connected PV systems of about 360 MW total capacity
were operational, of which 210 MW were installed in that year (IEA, 2013c). In
September 2012, projects with around 1.8 GW capacity had signed PPAs, projects
with 76 MW already had a letter of intent (LOI) and another 925 MW of projects
were waiting for an LOI (Kruangam, 2013).
626 A. Jäger-Waldau
In 2003, IDCOL started its Solar Energy Programme to promote the dissemina-
tion of solar home systems (SHSs) in the remote rural areas of Bangladesh, with
financial support from the World Bank, the Global Environment Facility (GEF),
the German Kreditanstalt für Wiederaufbau (KfW), the German Technical Coop-
eration (GTZ), the Asian Development Bank and the Islamic Development Bank.
By April 2013, the programme had seen more than 2 million SHSs installed in
Bangladesh (World Bank, 2013). At the time of writing, the installation rate is
more than 60,000 units per month. In 2011, the Asian Development Bank agreed
to provide financial support to Bangladesh for implementing the installation of
500 MW within the framework of the Asian Solar Energy Initiative (The Daily
Star, 2011; UNB Connect, 2011).
Indonesia
The development of renewable energy in Indonesia is regulated in the context of
national energy policy by Presidential Regulation No. 5/2006 (The President of the
Republic of Indonesia, 2006). This decree states that 11% of the national primary
energy mix in 2025 should come from renewable energy sources and the target
for solar PV was set at 1000 MW by 2025. By the end of 2011, about 20 MW of
solar PV systems had been installed, mainly for rural electrification purposes. In
2012, the Indonesian Ministry of Energy and Mineral Resources (ESDM) drafted
a Roadmap that foresaw installations of 220 MW between 2012 and 2015 (ESDM,
2012). A new policy to promote solar energy through auction mechanisms was
published in June 2013 (Bloomberg New Energy Finance, 2013a); how this new
policy will influence the market remains to be seen.
Kazakhstan
The development of renewable energy was one of the priorities of Kazakhstan’s
State Programme for Accelerated Industrial and Innovative Development for 2010–
2014. The main goal of this programme is to develop a new and viable economic
sector for growth, innovation and job creation. In addition, it drives the develop-
ment of renewable energy sources for the electricity sector in Kazakhstan and is
regulated by the Law on Support to the Use of Renewable Energy Sources, adopted
in 2009 (CIS Countries Legislation Database, 2009). In February 2013, the Kazakh
government decided to install at least 77 MW of PV by 2020 (Government of Kaza-
khstan, 2013). In 2011, JSC NAC Kazatomprom, jointly with a French consortium
headed by Commissariat à l’Energie Atomique et aux Energies Alternatives (CEA),
started the project Kaz PV with the aim of producing PV modules based on Kazakh
silicon (Kaz Silicon, 2011). The first project step was concluded in January 2013,
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch13 page 627
Malaysia
The Malaysia Building Integrated Photovoltaic (BIPV) Technology Application
Project was initiated in 2000, and by the end of 2009, a cumulative capacity of
about 1 MW of grid-connected PV systems had been installed under this pro-
gramme. The Malaysian government officially launched their GREEN Technology
Policy to encourage and promote the use of renewable energy for Malaysia’s future
sustainable development in 2009. By 2015, about 1 GW is planned to come from
renewable energy sources, according to the Ministry of Energy, Green Technology
and Water.
In April 2011, renewable energy FiTs were approved by the Malaysian Par-
liament, with a target of 1.25 GW installed by 2020. For the period December
2011–June 2014, a total quota of 125 MW was allocated for PV. The 2013 tariffs
were set by the Sustainable Energy Development Authority between 0.782 and
1.555 MRY/kWh (0.195 to 0.389 /kWh13 ), depending on the type and system
size. In addition, there is a small bonus for local module or inverter use. An annual
reduction in these tariffs is foreseen. As of 30 April 2013, 28.92 MW of PV systems
were operational under the new FiT scheme and another 141.58 MW had received
approval, and were in various stages of project planning or installation (SEDA,
2013). First Solar (USA), Hanwha Q Cells (Korea/Germany), Sunpower (USA),
and recently Panasonic (Japan), have set up manufacturing plants in Malaysia, with
more than 3.8 GW of total production capacity.
The Philippines
The Philippines’ Renewable Energy Law was passed in December 2008 (Philippine
Department of Energy, 2008). Under this law, the country had to double the energy
it derives from renewable energy sources within ten years. In June 2011, the Energy
Secretary unveiled a new Renewable Energy Roadmap, which aimed to increase
the share of renewables to 50% of total energy consumption by 2030, and to boost
renewable energy capacity from the 2011 level of 5.4 GW to 15.4 GW by 2030.
In early 2011, the country’s Energy Regulator National Renewable Energy Board
recommended a target of 100 MW of solar installations to be constructed in the
628 A. Jäger-Waldau
country over the following three years. An FiT of 17.95 PHP/kWh (0.299 /kWh)14
was suggested, to be paid from January 2012 onwards. For 2013 and 2014, an
annual degression of 6% was foreseen. The initial period of the programme is
scheduled to end on 31 December 2014. In July 2012, the Energy Regulatory
Commission decided to lower the tariff in view of lower system prices to 9.68
PHP/kWh (0.183 /kWh15 ) and confirmed the degression rate. By the end of 2012,
about 2 MW of the 20 MW of installed PV systems were grid-connected. SunPower
had two cell-manufacturing plants outside of Manila, but closed down Fab. No 1
in early 2012. Fab. No 2, with a nameplate capacity of 575 MW, remains in action.
Vietnam
The National Energy Development Strategy of Vietnam was approved in December
2007. This gave priority to the development of renewable energy and included the
following targets: increase the share of renewable energies from negligible to about
3% (58.6 GJ) of the total commercial primary energy in 2010, to 5% in 2020, 8%
(376.8 GJ) in 2025, and 11% (1.5 TJ) in 2050. By the end of 2011, about 5 MW
of PV systems had been installed, mainly in off-grid applications. In May 2011,
the Indo-Chinese Energy Company broke ground in the central coastal Province
of Quang Nam for the construction of a thin-film solar panel factory with an initial
capacity of 30 MW and a final capacity of 120 MW. However, the company applied
for permission to delay the project in June 2012, with no new launch date set. In
January 2013, WorldTech Transfer Investment and the UAE-based Global Sphere
began work on a solar-panel manufacturing plant in the central province of Thua
Thien-Hue (Global Sphere, 2013). This plant is located in the Phong Dien Industrial
Park and the first phase of the project (60 MW) should be operational by June 2015.
The capacity is planned to be increased to 250 MW in a second phase.
100000
2000
10000
2010
2012
1000
Cumulative installed [MW]
100
10
0.1
0.01
0.001
DE IT ES FR BE CZ UK EL BG SK AT DK NL PT SI LU SE MT CY FI RO HU PL LT LV IE EE HR TR
The net growth of all renewable energy power generation capacity in Europe
and Turkey between 2000 and 2012 was 178 GW. This compares with 121 GW
of new gas-fired capacity and a reduction in capacity of generation from coal
(−12.7 GW), oil (−7.4 GW) and nuclear (−14.7 GW). Wind (96.7 GW) and PV
(69 GW) accounted for more than 93% of the renewable capacity. This net growth
of 178 GW to a total of 316 GW of renewable capacity increased the total share of
renewable power capacity in Europe and Turkey from 22.5% in 2000 to 33.9% in
2012.
A total of about 45 GW of new power capacity was connected in the European
Union (EU) in 2012 and 12.5 GW was decommissioned, resulting in 32.5 GW
of new net capacity, as shown in Fig. 13.5 (European Wind Association, 2013;
Systèmes Solaires, 2013). Photovoltaic electricity generation capacity accounted
for 16.8 GW, or 51.7%, of the new net capacity. Wind power was second with
11.7 GW (36%), followed by 5 GW (15.4%) gas-fired power stations, 1.3 GW (4%)
biomass, 0.8 GW (2.5%) solar thermal power plants, 266 MW (0.8%) hydro and
61 MW (0.2%) other sources. The net installation capacity for coal-fired, oil-fired
and nuclear power plants was negative, with decreases of 2.3 GW, 3.2 GW and
1.2 GW, respectively. The renewable share of new power installations was more
than 69% and more than 95% of new net capacity in 2012.
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch13 page 630
630 A. Jäger-Waldau
Oil
Installed Capacity
Nuclear Decommissioned Capacity
others
Hydro
CSP
Biomass
Coal
Gas
Wind
PV
-6 -4 -2 0 2 4 6 8 10 12 14 16
New Installed or Decomissioned Capacity in 2012 [GW]
Belgium
Belgium showed another strong PV market performance in 2012, with new system
installations of 882 MW bringing the cumulative installed capacity to 2694 MW.
Photovoltaic power supplied 14% of Belgium’s residential consumption, or 2.8%
of the country’s total electricity needs. However, most of this capacity was again
installed in Flanders, where green certificates have been in place for 20 years.
The certificate system was suspended at the end of July 2012 and replaced by a
new regime for PV systems commissioned after 1 August 2012 (Belgisch Staats-
blad, 2012, 2013): for systems smaller than 250 kWp, the tariff was reduced from
0.23 /kWh to 0.21 /kWh, and for larger systems from 0.15 /kWh to 0.09 /kWh
for systems installed before the end of 2012. At the same time, the duration for
which the certificates could be claimed was reduced from 20 to 10 years.
Since 1 January 2013, the right to receive green certificates has been deter-
mined by the duration of the amortisation period — for PV it was 15 years
in 1H 2013 and the net metering scheme for systems below 10 kW continued.
A technology-dependent banding factor, which is set twice a year by the Flemish
Energy Agency, was introduced, so that one certificate is no longer for 1 MWh,
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch13 page 631
but now depends on the type of installation (RES LEGAL Europe, 2013). Since 1
January 2013 the value of the certificates has been 93.
In Wallonia, the green certificate scheme called SOLWATT was replaced by
a new scheme called QUALIWATT from 1 April 2013 (Government of Wallonia,
2013). The main change was that green certificates with a value of 0.065 /kWh
will only be granted until the time that the PV installation is fully reimbursed,
and their maximum term will be 10 years. In addition, for systems below 10 kW,
net metering continues and a progressive pricing system has been introduced. The
value of green certificates in the Brussels region is 65 and for PV systems there
is a multiplier of 2.2. In addition, the possibility of net metering exists for systems
below 5 kWp so long as the generated electricity does not exceed the in-house
electricity demand of the owner.
Bulgaria
A new Renewable Energy Source (RES) Act was approved in May 2011. This fixed
the FiT levels and resulted in new installations of around 110 MW, increasing the
total installed capacity to 134 MW by the end of 2011. At the end of 2012, 933 MW
of PV systems had been cumulatively installed. In March 2012, the Bulgarian
Parliament voted to revise the RES Act (Bulgarian State Gazette, 2012a). The
most significant change was that the price at which electricity would be purchased
was no longer fixed at the date of completion of installation, but at the date the
usage permit was granted.
In September 2012, the Bulgarian State Energy and Water Regulatory Com-
mission (SEWRC) published prices for the retroactive grid-usage fee, in accor-
dance with the Energy Act amendments adopted by the Bulgarian Parliament in
July 2012 (Bulgarian State Gazette, 2012b) for access to the transmission and dis-
tribution grid. For PV systems commissioned after 1 April 2010 and in 2011, the
fee amounted to 20% of the FiT. For systems commissioned during the first half
of 2012, the fee was 39% of the FiT, for those commissioned between 1 July and
31 August 2012 it was 5%, and after 1 September 2012, 1% of the respective FiT
applied (SEWRC, 2012a).
Modified FiTs, mandating a 34–54% reduction, depending on system type,
came into force on 1 June 2012. A further reduction ranging between 5 and 35%
came into force in October 2012 (SEWRC, 2012b).
Denmark
Due to the introduction of a net-metering system and the country’s high electricity
prices of 0.295 /kWh, 378 MW of PV systems were installed in Denmark in
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch13 page 632
632 A. Jäger-Waldau
2012. Following this rapid development, the regime was changed in November
2012 (Denmark, 2012). Under the new rules, full net metering was only possible
within one hour from when PV electricity was produced, and the excess electricity
exported to the grid was reimbursed at the rate of 1.30 DKK/kWh (0.174 /kWh16 )
in 2013. To account for decreasing PV system prices, this rate will decrease to 1.17
DKK/kWh (0.157 /kWh) in 2014 and to ∼1.00 DKK (0.174 /kWh) in 2015.
After ten years, the rate will be further reduced to about 0.60 DKK/kWh (0.080
/kWh).
France
In 2012, 1.08 GW of new PV systems were installed in France, which led to an
increase of the cumulative installed capacity to over 4 GW, including about 300 MW
in the French Overseas Departments. New PV installations in mainland France
accounted for 35% of total new electricity production capacity commissioned in
2012. Of the total capacity, residential systems smaller than 3 kWp represented 16%
or 0.64 GW, systems up to 250 kWp accounted for 40% or 1.6 GW, and systems
larger than 250 kWp added 44% or 1.76 GW.
In 2013, France had three different support schemes for PV. For systems up to
100 kWp, there is the FiT (allocation of 200 MW for residential and 200 MW for
commercial applications); for rooftop systems between 100 and 250 kWp there is
a ‘simplified’ call for tender with a volume of 120 MW for 2013; and for systems
larger than 250 kWp (large rooftop and ground-mounted systems) an additional
call for tender with a volume of 400 MW was issued, which received more than
1.9 GW of project applications.
In 2012, four PV tenders were offered and the average electricity sale price
proposed by the bidders fell from 229 /MWh during the first round to 194 /MWh
in the fourth one. New FiTs were published in February 2013, foreseeing an adap-
tation every three months (Ministère de l‘Écologie, 2013). Photovoltaic systems
with defined European content were eligible for a bonus of 5 to 10%.
Germany
Photovoltaics installed in Germany in 2012 increased slightly compared with 2011,
from 7.5 GW to 7.6 GW (Bundesnetzagentur, 2013). German market growth is
directly correlated to the introduction of the Erneuerbare Energien Gesetz (Renew-
able Energy Sources Act) in 2000 (EEG, 2000). This introduced a guaranteed FiT
for electricity generated from solar PV systems for 20 years and had a built-in
annual decrease, which was adjusted over time to reflect the rapid growth of the
market and the corresponding price reductions. As only estimates of the installed
capacity existed before 2009, a plant registrar was introduced from 1 January 2009
onwards.
The German market performed strongly throughout 2012 with peaks of 1.2 GW
installed in March, 1.8 GW in June and 1 GW in September. The total installed
capacity at the end of the year was 32.7 GW. Since May 2012, the FiT has been
adjusted on a monthly basis depending on the actual installations during the
previous quarter. The fact that the tariff for residential PV systems (July 2013:
0.151 /kWh) is now below the rate conventional electricity consumers are pay-
ing (0.287 /kWh) makes the increase of self-consumption more attractive and is
opening new possibilities for the introduction of local storage (Bundesnetzagentur,
2013; Strategies Unlimited, 2013). Since 1 May 2013, the Kreditanstalt für Wieder-
aufbau (KfW) has offered low interest loans with a single repayment bonus of up
to 30% and a maximum of 600 per kW of storage for PV systems up to 30 kWp
(KfW, 2013). The maximum repayment bonus is limited to 3000 per system.
Greece
In 2009, Greece introduced a generous FiT scheme but this had a slow start until the
market accelerated in 2011 and 2012. In 2012, 687 MW of new PV systems were
installed — more than 1.5 times the 439 MW that had been cumulatively installed
up to the end of 2011. In April 2013, the total installed capacity surpassed 2 GW
(over 1.9 GW in mainland Greece and over 115 MW in the Islands) (HEDNO S.A.,
2013; LAGIE, 2013). On 10 May 2013 the Greek Ministry of Environment, Energy
and Climate Change announced retroactive changes in the FiT for systems larger
than 100 kWp and new tariffs for all systems from 1 June 2013 onwards.
Italy
Italy connected more than 3.5 GW during 2012, bringing cumulative installed
capacity to 16.4 GW by the end of the year (Gestore Servizi Energetici, 2013a).
The Quinto Conto Energia (Fifth Energy Bill) was approved by the Italian Council
of Ministers on 5 July 2012 (Italy, 2012). This set new half-yearly reductions of the
tariffs, and the annual expenditure ceiling for new installations was increased from
500 million to 700 million. In addition, a new requirement to register systems
larger than 12 kWp was introduced. On 6 June 2013, Gestore Servizi Energetici
announced that the 6.7 billion ceiling of the bill had been reached with 18.2 GW,
out of which 17.1 GW were already operational, and that the Quinto Conto Energia
would cease within 30 days, as foreseen (Gestore Servizi Energetici, 2013b).
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch13 page 634
634 A. Jäger-Waldau
Slovakia
After two years of rapid growth, the Slovakian market decreased by over 90% in
2012 and only 29 MW were newly installed. The total capacity of 517 MW is more
than three times the original 160 MW capacity target for 2020, published in the
National Renewable Energy Action Plan in 2010. From February 2011, support
was limited to applications for systems smaller than 100 kW, and from 1 July 2013
onward, support was limited to systems of up to 30 kW placed on buildings.
Spain
Spain is still third in Europe for cumulative PV capacity, with 4.5 GW installed
by the end of 2012. Most of this was installed in 2008, when the country was
the biggest market, with close to 2.7 GW in 2008 (European Photovoltaic Industry
Association, 2013). This was more than twice the expected capacity and was due to
an exceptional race to install systems before the Spanish Government introduced
a cap of 500 MW on the yearly installations in the autumn of 2008. Royal Decree
1758/2008 set considerably lower FiTs for new systems and limited the annual
market to 500 MW, with the proviso that two-thirds should be rooftop-mounted
and not free-field systems. These changes resulted in a steep drop in the number
of new installations. In 2012, new system installations with a capacity of 194 MW
increased total capacity to 4.5 GW. Photovoltaic-generated electricity contributed
7.8 TW, or 2.9%, of the Spanish power demand in 2012.
In January 2012, the Spanish government passed a further Royal Decree 1/12
(Government Gazette, 2012), which suspended the remuneration pre-assignment
procedures for new renewable power capacity, affecting about 550 MW of planned
solar PV installations. The justification given for this move was that Spain’s energy
system had by then amassed a 30 billion power-tariff deficit and it was argued
that the special regime for renewable energy was the main reason. However, the
Spanish government has prevented utilities from charging consumers for the true
cost of electricity for over a decade. Instead of allowing utilities to increase rates
every time electricity generation costs increased (due to rising coal or natural gas
costs, inflation or changes in energy or environmental policy), the government
allowed them to create schemes such as a deferral account whereby they could
recover shortfalls in any individual year from revenues generated in subsequent
years.
In January 2007, the European Commission opened an in-depth investigation
to examine the potential support to large and medium-sized companies and elec-
tricity distributors in Spain, in the form of artificially low regulated industrial tariffs
for electricity (EUR-Lex, 2007). In 2005, these regulated tariffs led to a deficit of
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch13 page 635
3.8 billion in the Spanish electricity system, which increased to almost 9 billion
in 2007, a time when payments under the special regime for renewable energy were
still limited.
Despite the Royal Decree 1/12, the tariff deficit increased further in 2012 and
reached 35.6 billion at the end of February 2013 (Comición National de Energía
2013). The question remains: is renewable electricity generation responsible for
this? Despite the fact that the premium payments amount to 8.4 billion in 2012
and 9.1 billion for 2013, the answer is not clear. An often-neglected aspect is the
fact that investment in renewable capacity has increased supply at the wholesale
market, thus decreasing the system marginal price by the order of merit effect. It is
argued that renewable energy ‘pays for itself’ because by bidding at the pool at
zero prices these units substantially decrease the system marginal price. Therefore
the cost of all electricity should be lower, if this price reduction is passed on to
the customer. An analysis for the entire special regime concluded that for 2010 the
decrease in the wholesale price of around 29 /MWh covered 70% of the FiT costs
(Ciarreta et al., 2012b).
Switzerland
About 200 MW of PV systems were installed in Switzerland in 2012, almost dou-
bling the total capacity to 411 MW. Prices for turnkey systems decreased by over
40% in 2012 (Photovoltaik Guide, 2013). In view of these decreasing prices, the
FiT was reduced three times in 2012.
United Kingdom
The United Kingdom introduced a new FiT scheme in 2010, which led to the
installation of approximately 55 MW that year and over 1 GW in 2012. This steep
increase was caused by the announcement of a fast-track review of large-scale
projects by the Department of Energy & Climate Change (DECC) in February
2011, which led to a rush to complete these projects during the first half of 2011
(DECC, 2011). A second rush occurred towards the end of the year to meet the
deadline of 12 December 2011, when the DECC planned to decrease the residential
tariff by about 50% as a result of another fast-track consultation. However, this
decision was contested in court and the tariffs were only changed on 1 April 2012.
The average reductions in April were 44–54% for systems smaller than 50 kWp
and 0–32% for systems above 50 kWp. In November 2012, a further reduction
of 3.5% for systems smaller than 50 kWp was imposed. No reduction was made
for larger systems, as almost no such systems had been installed between May
and July 2012. However, for larger systems it was possible to receive Renewables
Obligation Certificates (ROCs) until April 2013; for PV the rate was two ROCs
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch13 page 636
636 A. Jäger-Waldau
per MWh. Overall about 630 MW were installed, bringing the cumulative capacity
to 1.8 GW at the end of 2012.
During 2012, the Energy and Climate Change Minister, Greg Barker, repeat-
edly declared his desire to see the UK solar market reach 22 GW by 2020. During
the first six months of 2013, the UK added 802 MW of new solar PV installations,
representing the strongest first-half year ever for the UK PV industry. This was
comprised of 20 MW in the first quarter of 2013, followed by 282 MW in the
second quarter of 2013 (Colville, 2013).
Austria
In 2012 Austria installed about 230 W of new PV systems and more than doubled
its cumulative capacity to 417 MW. The Ökostrom-Eispeisetarifverordnung 2012
(eco-electricity decree) is the regulation that sets the prices for the purchase of
electricity generated by green power plants. An investment subsidy with a budget
of 36 million in 2013 is also in place. Regardless of the size of the systems, a
maximum of 5 kWp is supported, with 300 /kWp for add-on and ground-mounted
systems and 400 /kWp for building-integrated systems. Operators of PV systems
larger than 5 kWp can choose to opt for the so-called net-parity tariff (Netzparitäts-
Tarif) for a period of 13 years. Since 1 January 2013, this option has only been
available for systems on buildings.
Portugal
Despite high solar radiation, solar photovoltaic system installation in Portugal has
grown only very slowly, reaching a cumulative capacity of 229 MW at the end of
2012.
Ukraine
Ukraine again saw an impressive growth and almost doubled its capacity with over
180 MW new installed PV systems to 370 MW. In February 2013, the Ukrainian
Parliament had the first reading of a bill to simplify the access of households to
the feed-in scheme. Reduction in the feed-in tariffs by between 16% and 27%,
depending on the type of installation, was finally adopted in February 2013 and
went into force on 1 April 2013. Since then, tariffs have been adjusted on a monthly
basis. In 2013 another doubling of the installed capacity is possible.
Turkey
In March 2010, the Energy Ministry of Turkey unveiled its 2010–2014 Strategic
Energy Plan. One of the government’s priorities was to increase the proportion
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch13 page 637
13.2.4 Africa
Despite its vast solar resources and the fact that there are areas in Africa where
the solar potential is very considerable — with the same photovoltaic panel able to
produce twice as much electricity in Africa as in Central Europe on average — only
limited use of PV electricity generation is made in Africa. The main application
is for small solar home systems, and the market statistics for these are extremely
imprecise or non-existent. Therefore, all African countries are considered potential
or emerging markets and some of them are mentioned below.
Algeria
In 2011, the Algerian Ministry of Energy and Mines published its Renewable
Energy and Energy Efficiency Programme, which aimed to increase the share of
renewable energy used for electricity generation to 40% of domestic demand by
2030. The plan foresaw 800 MW of installations by 2020 and a total of 1.8 GW by
2030. It is estimated that about 5 MW of small decentralised systems were installed
at the end of 2012. For 2013, new installations of around 20 MW were planned.
Capo Verde
Capo Verde’s Renewable Energy Plan (2010–2020) aims to increase the use of
renewable energy to 50% of total energy consumption in 2020. The policy to
achieve this is to use PPAs. Law n1/2011 established the regulations regarding
independent energy production. In particular, it established the framework con-
ditions for the set-up of independent power producers using renewable energy
(15 years PPA), and for self-production at user level. It created the micro-generation
regime, regulated rural electrification projects and established the tax exemption
of all imported renewable energy (RE) equipment. At the end of 2012, 7.5 MW
of centralised grid-connected PV systems were installed. In addition, there are a
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch13 page 638
638 A. Jäger-Waldau
number of smaller off-grid and grid-connected systems. To realise the 2020 50%
renewable energy target, about 340 MW of PV systems will be required.
Ghana
In 2011 the Parliament of Ghana passed a Renewable Energy Bill with the aims of
increasing the proportion of renewable energy, particularly solar, wind, mini hydro
and waste-to-energy schemes, in the national energy supply mix, and contributing
to the mitigation of climate change (Parliament of Ghana, 2011). The bill set a goal
of renewable energy constituting 10% of national energy generation by 2020. By
the end of 2012, a few thousand solar home systems and a few of off-grid systems
with an estimated 5 MW had been installed in the country. In 2012, Episolar,
Canada, signed a PPA with Ghana’s second largest utility, the Electricity Company
of Ghana, for a 50 MW PV plant with an option to increase the overall project size
to 150 MW (Episolar, 2012). In December 2012, Blue Energy, UK, announced that
it had agreed with Ghana’s Public Utilities Regulatory Commission on a 20-year
PPA for the 155 MW PV plant in Nzema to be operational by 2015 (Blue Energy,
2012). These two projects are still in the development phase looking for strategic
investors. In May 2013, the Volta River Authority inaugurated its first solar power
plant at Navrongo with a capacity of 2 MW. The Volta River Authority plans to
install a total of 14 MW by 2014.
Kenya
In 2008, Kenya introduced FiTs for electricity from renewable energy sources, but
solar was only included in the revision dated 2010 (Ministry of Energy, Kenya,
2010). However, only a little over 560 kW of PV capacity was connected to the
grid in 2011. The majority of the 14 MW of PV systems were off-grid installations.
In 2011, Ubbink East Africa Ltd., a subsidiary of Ubbink B.V. (Doesburg, The
Netherlands) opened a solar PV manufacturing facility with an annual output of
30,000 modules in Naivasha, Kenya. This plant produces modules for smaller PV
systems such as solar home systems. The estimates for the PV market in Kenya are
average sales of 20,000–30,000 home systems and 80,000 solar lanterns per year.
Morocco
The Kingdom of Morocco’s solar plan was introduced in November 2009, with the
aim of establishing 2000 MW of solar power by 2020. To implement this plan, the
Moroccan Agency for Solar Energy was founded in 2010. Both solar electricity
technologies, concentrated solar thermal power (CSP) and PV, were to compete
openly. In 2007, the National Office of Electricity announced a smaller programme
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch13 page 639
South Africa
South Africa has a fast-increasing electricity demand and vast solar resources. In
2008, the country enacted its National Energy Act, which called for a diversifica-
tion of energy sources, including renewables, as well as fuel switching to improve
energy efficiency (South Africa, 2008). In 2011, the Renewable Energy Indepen-
dent Power Producer Procurement Programme (IPP) was set up with rolling bidding
rounds. Two rounds took place in 2011 and 2012, and a third one closed in August
2013. The overall target is 3.725 GW, including 1.45 GW for solar PV. In the first
two bidding rounds 1048 MW of solar PV projects were allocated to preferred
bidders. The average bid price decreased between the first round (closure date: 4
November 2011) from 2.65 ZAR/kWh (0.252 /kWh17) to 1.65 ZAR/kWh (0.157
/kWh) in the second round (closure date: 5 March 2012). It is estimated that about
30–40 MW of PV systems were installed in South Africa by the end of 2012.
640 A. Jäger-Waldau
systems larger 500 kW would no longer be eligible for the FiT. Following the WTO
ruling in May on the local content regulations, the Minister said that Ontario would
comply with it.
Argentina
In 2006, Argentina passed its Electric Energy Law, which established that 8% of
electricity demand should be generated by renewable sources by 2016 (Argentina,
2006). It also introduced FiTs for wind, biomass, small-scale hydro, tidal, geother-
mal and solar for a period of 15 years. In July 2010, amongst other renewable
energy sources, the government awarded PPAs to six solar PV projects totalling
20 MW. By the end of 2012, about 17 MW (7 MW grid-connected) of PV systems
had been installed. According to the renewable energy country attractiveness indi-
cator, the Argentinean government has set a 3.3 GW target for PV installations by
2020 (Ernst & Young, 2011).
Brazil
At the end of 2012, Brazil had about 20 MW cumulative installed capacity, mainly
in rural areas. In April 2012, the board of the National Agency of Electric Energy
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch13 page 641
Chile
In February 2012, the President of Chile announced a strategic energy plan to
reach 20% of non-conventional renewable energy by 2020. In the first quarter of
2012, the first MW-size PV system was installed in the northern Atacam desert. By
June 2013, Chile’s Environmental Assessment Service had approved over 4 GW
of projects, with an additional 2.2 GW still under review. At the end of 2012, it
was estimated that about 10 MW of PV systems (5 MW grid-connected) had been
installed. In June 2013, about 70 MW of PV systems were under construction and
it is expected that about 50 MW of new systems will be operational at the end of
2013.
Dominican Republic
In 2007, the Dominican Republic passed a law promoting the use of renewable
energy and setting a target of 25% renewable energy share in 2025 (Domini-
can Republic, 2007). At that time ∼1–2 MW of solar PV systems had been
installed in rural areas, which had increased to over 5 MW by 2011. In 2011,
the first PPA for 54 MW was signed between Grupo de Empresas Dominicanas de
Energía Renovable and Corporación Dominicana de Empresas Eléctricas Estatales
(CDEEE). The first phase (200 kW) of the project became operational in July 2012,
and the whole solar farm was connected to the grid early 2013. In 2012, CDEEE
signed two more PPAs with a total capacity of 116 MW.
Mexico
In 2008, Mexico enacted a Law for Renewable Energy Use and Financing Energy
Transition to promote the use of renewable energy (Mexico, 2008). In 2012, the
country passed its Climate Change Law, which foresaw a decrease in greenhouse
gas emissions of 30% below the business-as-usual case by 2020, and 50% by 2050
(Mexico, 2012). It further stipulated that renewable electricity should constitute
35% of total supply by 2024. By the end of 2012, about 52 MW of PV systems had
been installed (IEA, 2013c).
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch13 page 642
642 A. Jäger-Waldau
Peru
In 2008, Peru passed its Legislative Decree 1002, which made the development
of renewable energy resources a national priority. The decree stated that at least
5% of electricity should be supplied from renewable sources by 2013. In February
2010, the Organismo Supervisor de la Inversión en Energía y Minería held the first
round of bidding and awarded four solar projects with a total capacity of 80 MW.
A second bidding round was held in 2011, with a quota of 24 MW for PV. By
the end of 2012, about 85 MW of PV systems had been installed. A doubling of
capacity is regarded as possible in 2013.
644 A. Jäger-Waldau
18 High concentration > 300 Suns (HCPV), medium concentration 5 < x < 300 Suns (MCPV), low
concentration < 5 Suns (LCPV).
19 Solar cell production capacities mean: in the case of wafer silicon-based solar cells, the cell; in the
case of thin-films, the complete integrated module. Only those companies that actually produce the
active circuit (solar cell) are counted. Companies that purchase these circuits and make cells are not
counted.
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch13 page 645
646 A. Jäger-Waldau
648 A. Jäger-Waldau
given. This must, however, be lower, because the company states in its financial
reports that it buys cells from other manufacturers. External reports gave cell pro-
duction as 1.24 GW (PV News, 2013).
650 A. Jäger-Waldau
in R&D, as well as the manufacture of solar cells, modules, equipment parts and
devices which exploit heat. In 2012, Kyocera had a production of 800 MW and
planned to increase its capacity to 1 GW in 2013 (Ikki, 2013ab).
cells and modules, was installed in May 2007 and an increase in production capacity
to 100 MW was achieved in July 2008. Commercial production of micromorph
solar modules started in July 2009. Thin-film capacity was 30 MW in 2010 and
increased to 75 MW in early 2011. Total production capacity for 2012 was reported
as 800 MW, with 584 MW of production (ENF, 2013).
652 A. Jäger-Waldau
for PV modules. This extreme price hike triggered a massive capacity expansion,
not only by established companies, but attracting many new entrants as well.
These massive production expansions, as well as the difficult global economic
situation, led to price decreases for polysilicon throughout 2009. Prices reduced to
about 50–55 $/kg at the end of 2009, with a slight upwards tendency throughout
2010 and early 2011, before they dropped significantly again. The lowest level on
the spot market was reached in December 2012 with prices below 16 $/kg (12.3
/kg) before they started to recover slightly, trading between 16.5 and 17.5 $/kg
(12.7–13.5 /kg) for most of 2013.
China was and remains a major polysilicon producer. In January 2011, the
Chinese Ministry of Industry and Information Technology tightened the rules for
future polysilicon factories to combat the oversupply problem. New factories must
be able to produce more than 3000 metric tons of polysilicon per year and also
meet certain efficiency, environmental and financing standards. Their total energy
consumption must be less than 200 kWh/kg of polysilicon, and China is aiming to
create large companies with at least 50000 metric tons annual capacity by 2015.
These two framing conditions, in addition to the enormous price pressure, are the
reasons why a significant number of Chinese manufacturers closed down their
polysilicon production in the first half of 2012. This is also why China imported
83000 metric tons of silicon in 2012, 32% more than 2011 (Bloomberg New Energy
Finance, 2013g).
Projected global silicon production capacities in 2013 vary between 290,000
metric tons (Bloomberg New Energy Finance, 2013e) and 409,690 metric tons
(Ikki, 2013a). It is estimated that about 27000 metric tons will be used by the
electronics industry. The possible solar cell production will, in addition, depend
on the material used per Wp; the current worldwide average is about 5.6 g/Wp.
ton factory in Nünchritz Saxony, started production in 2011. In 2010, the company
decided to build a polysilicon plant in Tennessee with 15000 ton capacity. The
groundbreaking of this new 18000 metric ton factory took place in April 2011,
and construction should be finished by the end of 2013. In addition, the company
also expanded its Burghausen capacity by 5000 metric tons in 2012 and, together
with a further expansion of the Nünchritz factory by 5000 metric tons, it plans to
have 70000 metric tons of production capacity in 2014. Total polysilicon sales are
reported as 38000 metric tons in 2012.
654 A. Jäger-Waldau
OCI has also invested in downstream business and holds 89.1% of OCI Solar
Power, which develops, owns and operates solar power plants in North America.
On 23 July 2012, the company signed a PPA with CSP Energy, Texas, for a 400 MW
solar farm in San Antonio, TX.
656 A. Jäger-Waldau
2012, expansion Phase 2 with 3000 metric tons came online, and according to the
company it will reach a total production capacity of 9300 metric tons at the end
of 2013. The company invested in the downstream business, ranging from wafers,
cells, modules and projects. At the end of 2011, it had a manufacturing capacity of
125 MW wafers and 100 MW modules. Polysilicon shipments of 3585 metric tons
were reported for 2012.
13.4 Outlook
The photovoltaic industry has changed dramatically in the last few years. China has
become the major manufacturing country for solar cells and modules, followed by
Taiwan, Germany and Japan. Amongst the 20 biggest photovoltaic manufacturers
in 2012, only three had production facilities in Europe, namely First Solar (USA,
Germany, Malaysia), Hanwha Q Cells (Germany and Malaysia) and REC (Norway
and Singapore). However, REC closed down its production in Norway in early 2012
and First Solar closed its factory in Germany later in the year.
It is important to remember that the PV industry consists of more than just
the cell and module production, and that the whole PV value chain is relevant.
Besides the information in this chapter, the whole upstream industry (e.g. materials,
equipment manufacturing), as well as the downstream industry (e.g. inverters, BOS
components, system development, installations) has to be looked at as well.
The implementation of the 100,000 roofs programme in Germany in 1990,
and the Japanese long-term strategy set in 1994, with a 2010 horizon, were the
beginning of an extraordinary PV market growth. Before the start of the Japanese
market implementation programme in 1997, annual growth rates of the PV markets
were in the range of 10%, mainly driven by communication, industrial and stand-
alone systems. Since 1990, PV production has increased by almost three orders
of magnitude, from 46 MW to about 38.5 GW in 2012. Statistically documented
cumulative installations worldwide accounted for almost 100 GW in 2012. This
represents mostly the grid-connected photovoltaic market. To what extent the off-
grid and consumer product markets are included is not clear, but it is believed that
a substantial part of these markets are not included in these figures as it is very
difficult to track them.
Even with the current economic difficulties, overall installations are increasing,
due to overall rising energy prices and the pressure to stabilise the climate. In the
long term, growth rates for photovoltaics will continue to be high, even if economic
conditions vary and lead to short-term slowdowns.
This view is shared by an increasing number of financial institutions, which
are turning towards renewables as a sustainable and stable long-term investment.
Increasing demand for energy is pushing the prices for fossil energy resources
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch13 page 657
higher and higher. In 2007, a number of analysts predicted that oil prices could
well hit 100 $/bbl by the end of that year or early 2008 (Bloomberg, 2007). After
the spike of oil prices in July 2008, at close to 150 $/bbl, prices decreased due to
the worldwide financial crisis and hit a low of around 37$/bbl in December 2008.
Since then, the oil price has rebounded and the IEA reported average prices for
oil imports around 110 $/bbl since the second quarter of 2011, with peaks around
120 $/bbl between February and April 2012 and September 2013. Oil demand has
increased from about 84 million bbl/day in Q1 2009 to around 91 million bbl/day
in Q3 2013, and a moderate further increase to 93 million bbl/day by Q4 2014 is
expected. Even though no significant changes are currently forecast by analysts
for 2014, the fundamental trend, that increasing demand for oil will drive its price
higher again, is intact and will be evident as soon as the global economy recovers.
Over the last 20 years, numerous studies of the potential growth of the PV
industry and the implementation of PV electricity generation systems have been
produced. In 1996, the Directorate-General for Energy of the European Commis-
sion (EC) published a study entitled Photovoltaics in 2010 (European Commission,
1996). The medium scenario of this study was used to formulate the EC’s White
Paper target of 1997, which was to have a cumulative installed capacity of 3 GW
in the EU by 2010 (European Commission, 1997). The most aggressive scenario in
this report predicted a cumulative installed PV capacity of 27.3 GW worldwide and
8.7 GW in the EU for 2010. This scenario was called ‘Extreme scenario’ and it was
assumed that a number of breakthroughs in technology and costs, as well as con-
tinuous market stimulation and elimination of market barriers, would be required
to achieve it. The reality check reveals that even the most aggressive scenario is
lower than what we expect from the current developments.
According to investment analysts and industry prognoses, solar energy systems
will continue to grow at high rates in the coming years. The different photovoltaic
industry associations, as well as Greenpeace, the European Renewable Energy
Council (EREC) and the IEA, have developed new scenarios for the future growth
of PV. Table 13.1 shows the various scenarios of the Greenpeace/EREC study,
as well as the different 2013 IEA World Energy Outlook scenarios and the IEA
PV Technology Roadmap of 2010. It is interesting to note that the 2015 capacity
values of only two scenarios — the Greenpeace (revolution) and IEA 450 ppm
Scenarios — were not reached at the end of 2013. With forecast new installations
of between 93 and 106 GW in 2014 and 2015, even the Greenpeace revolution
scenario is no longer fictional thinking (Bloomberg New Energy Finance, 2014).
The above-mentioned solar photovoltaic scenarios will only be possible if
solar cell and module manufacturing are continuously improved and novel design
concepts are realised, as with current technology the demand for some materials,
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch13 page 658
658 A. Jäger-Waldau
Table 13.1 Predicted cumulative solar electrical capacities until 2035. Data sources:
Bloomberg New Energy Finance, 2014; Greenpeace, 2012; IEA, 2010, 2013d.
∗ 2035 values are extrapolated, as only 2030 and 2040 values are given
∗∗ 2015 and 2030 values are extrapolated, as only 2011, 2020 and 2035 values are given
∗∗∗ 2015 and 2035 values are extrapolated, as only 2010, 2020, 2030 and 2040 values are given
like silver, would dramatically increase the costs for these resources within the
next 30 years. Research to avoid such problems is underway and it can be expected
that these bottle-necks will be avoided. With 100 GW cumulative installed pho-
tovoltaic electricity generation capacity installed worldwide by the end of 2012,
photovoltaics still is a small contributor to the electricity supply, but its importance
for our future energy mix is finally acknowledged.
References
Argentina (2006), Boletín Oficial de la República de Argentina, Número 31.064
(www.diputados-catamarca.gov.ar/ley/BO020107.pdf).
Belgisch Staatsblad (2012), 182e Jaargang, No 237, 20 Juli 2012, Derde Editie, N.
2012—2122 [C − 2012/35855] 13 JULI 2012.—Decreet houdende wijziging van het
Energiedecreet van 8 mei 2009, wat betreft de milieuvriendelijke energieproductie (1),
p. 40448ff.
Belgisch Staatsblad (2013), 183e Jaargang, No 11, 16 Januari 2013, [C − 2013/11001],
21 DECEMBER 2012.—Koninklijk besluit tot wijziging van het koninklijk besluit
van 16 juli 2002 betreffende de instelling van mechanismen voor de bevordering van
elektriciteit opgewekt uit hernieuwbare energiebronnen, p. 1574ff.
Bloomberg (2007), ‘$100 oil price may be months away, say CIBC, Goldman’, 23 July 2007
(http://www.bloomberg.com/apps/news?pid=newsarchive&sid=ajxtV4oWcHk0).
September 17, 2014 11:4 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-ch13 page 659
660 A. Jäger-Waldau
662 A. Jäger-Waldau
APPENDIX I
FUNDAMENTAL CONSTANTS
663
September 17, 2014 11:7 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-app page 664
APPENDIX II
USEFUL QUANTITIES
AND CONVERSION FACTORS
664
September 17, 2014 11:7 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-app page 665
APPENDIX III
LIST OF SYMBOLS
Preferred
Symbol Unit Description
a absorptance (absorptivity)
A electron acceptor
Ae eV electron affinity
Ah eV hole affinity
ci mol dm−3 molar concentration of species i
D electron donor
D(E) cm−3 eV−1 density of electronic states of energy E
Di cm2 s−1 diffusion coefficient of species i
De,h cm2 s−1 diffusion coefficient of electron or hole
E eV internal energy
molecule−1
or kJ mol−1
Ec , Ev eV conduction or valence band-edge energy
Eg eV bandgap energy or minimum energy for electronic
molecule−1 excitation
or kJ mol−1
EF eV Fermi level
E Fn , E F p eV Fermi level in n or p material
E Fe , E Fh eV quasi-Fermi level of an electron or hole
g cm−3 s−1 or volume rate of generation of excited states
m−3 s−1
G eV Gibbs free energy
molecule−1
or kJ mol−1
H eV enthalpy
molecule−1
or kJ mol−1
H m cell width
H p,n m width of the p or n layer
HRP eV electronic coupling matrix element between reactant (R)
molecule−1 and product (P) states
i mA cm−2 current density
i ph mA cm−2 photocurrent density
ij mA cm−2 junction (dark) current density
665
September 17, 2014 11:7 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-app page 666
666 Appendices
Preferred
Symbol Unit Description
Appendices 667
Preferred
Symbol Unit Description
p atm or Pa pressure
p cm−3 or hole concentration in the valence band of a semiconductor
m−3
po cm−3 or equilibrium hole concentration
m−3
P W cm−2 or power density
W m−2
q C charge of an electron
Q C electrical charge
Q p , Qn m width of the p- or n-layer quasineutral region
r cm−3 s−1 or volume rate of recombination of excited
m−3 s−1
rλ 1 reflectivity at wavelength λ
R resistance
R square−1 sheet resistance
RL load resistance
Se , Sh m s−1 surface recombination velocity for an electron or hole
T K absolute temperature
T transmittance (transmissivity)
ue , uh cm2 s−1 electron or hole mobility
V−1
uE cm s−1 drift speed in electric field strength E
UOx,Rd V vs. ref. electrode potential for the reaction Ox + ne− Rd vs.
reference electrode
Uo V vs. ref standard electrode potential (based on activities)
U o V vs. ref formal electrode potential (based on concentrations)
V m3 volume
V V voltage
Vb V band-bending potential
Vbo V equilibrium band-bending potential
Vt h V thermal voltage (β) kT/q (= 0.02569 V for β = 1 and
25◦ C)
α cm−1 or optical absorption coefficient
m−1
αλ cm−1 or optical absorption coefficient at wavelength
m−1
β diode quality factor
ε emissivity
o permittivity of a vacuum
op optical dielectric constant (op = n r2 )
s static dielectric constant
E V m−1 electric field strength
September 17, 2014 11:7 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-app page 668
668 Appendices
Preferred
Symbol Unit Description
APPENDIX IV
AM Air Mass
BHJ bulk heterojunction
BIPV Building Integrated Photovoltaics
BOS Balance Of System
BSF Back-Surface Field
cb (as a superscript or subscript) conduction band
CBE Chemical Beam Epitaxy
CLEFT Cleavage of Lateral Epitaxial Films for Transfer
CPV Concentrated Photovoltaics
CSP Concentrated Solar Power
CVD Chemical Vapour Deposition
CZ Czochralski (method of crystal growth)
DN(I) Direct Normal (Irradiance)
Dk (as a superscript or subscript) in the dark
DoS Density of States
EFG Edge-defined Film-fed Growth
EPC Engineering, Procurement and Construction
EQE External Quantum Efficiency
fb (as a superscript or subscript) flat band
FZ Floatzone (method of crystal growth)
G Global (irradiance)
HCPV High-Concentration Photovoltaics
IQE Internal Quantum Efficiency
ITO indium tin oxide
Lt (as a superscript or subscript) in the light
LCL lateral conduction layer
LGBC Laser-Grooved Buried Contact
LOLP Loss-Of-Load Probability
LPE Liquid Phase Epitaxy
MBE Molecular Beam Epitaxy
MIS Metal|Insulator|Semiconductor (junction or device)
MOCVD Metal Organic Chemical Vapour Deposition
mp (as a superscript or subscript) maximum power
MS Metal|Semiconductor (junction or device)
NOCT Nominal Operating Cell Temperature
oc (as a superscript or subscript) open circuit
669
September 17, 2014 11:7 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-app page 670
670 Appendices
Ox oxidant
PCE power conversion efficiency
PEDOT:PSS poly(3,4-ethylenedioxythiophene):poly(styrenesulphonate)
PTC Practical Test Conditions (for cell efficiency measurements):
(plane-of-array irradiance 1000 W m−2 AM1.5G, ambient temperature
25◦ C, wind speed air 1 m s−1 )
PPV poly( p-phenylenevinylene)
PV photovoltaic
QNR quasi-neutral region
RAPS Remote Area Power Supplies
Rd Reductant
sc (as a superscript or subscript) short circuit
sc1 (as a superscript or subscript) space-charge layer
SHE Standard Hydrogen Electrode
SHS Solar Home System
SIS Semiconductor|Insulator|Semiconductor junction
ss (as a superscript or subscript) surface state
STC Standard Test Conditions (for cell efficiency measurements):
plane-of-array irradiance AM1.5G 1000 W m−2 , cell temperature
25 ± 1/2 C
TCO Transparent Conducting Oxide
TPV thermophotovoltaic
vb (as a subscript or superscript) valence band
Wp peak watts
September 17, 2014 11:9 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-index page 671
INDEX
671
September 17, 2014 11:9 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-index page 672
672 Index
Index 673
Czochralski (CZ) crystal growth process, energy payback time, 144, 238, 239
90, 94, 106–109 Entech Solar, 539, 541
CZTS (kesterite) solar cells, 294 environmental impact of PV, 608
etendue, 46, 49, 51, 510, 511, 518, 519
Daqo New Energy, 655 Euro-efficiency value, 591
declination–hour angle tracking, 495, 498, EuroCIS, 16, 268
532 European Union
defect PV market, 618
charged, 151, 152, 156 excitons, 344
dangling bond, 148, 152 CT states, 375
density, 150, 152 exciton quenching, 345, 349, 394
equilibrium, 150–153 field-assisted dissociation, 374
neutral, 149, 151, 152 hot excitons, 376
Denmark triplet excitons, 395
PV market, 631 external quantum efficiency (EQE), 61,
Dexter energy transfer, 346 62, 274, 350
Doha Climate Change Conference, 6 a-Si cell, 156
doping, 151, 155 triple cell, 162
double-hetero (DH) structures, 309, 322, external radiative efficiency (ERE), 61, 65,
324 66
down-conversion, 74, 75, 81, 105
DSSC Förster energy transfer, 346
see dye-sensitised solar cells feed-in tariffs (FiTs), 2, 545, 620–622,
dye-sensitised solar cells (DSSCs) 625, 636
charge screening in, 419, 420 First Solar, 15, 81, 212, 627, 645, 656
device configuration, 415–417 floatzone (FZ) crystal growth process, 109
DSSC/CIGS tandem cell, 432 fluorene, 387
device fabrication, 417, 419 fluorenone, 387
dye regeneration, 419–421, 424 France
energetics of operation, 418–422 PV market, 632
historical background, 414 free carrier absorption (FCA), 176, 185
kinetics of operation, 422–426 fullerenes, 352, 353, 358, 360, 361, 364,
molecular engineering of interface, 370–372, 374–377, 383, 384, 389, 393,
439, 440 395, 400
organic dyes in, 436, 437
panchromatic sensitisers, 426–429 GaAs, 10, 11, 27, 29, 31, 33
quantum dots as sensitisers, 437, 438 cell history, 16
solid-state DSSC, 440, 441 in QW solar cells, 456
stability, 433–436 minority-carrier lifetimes, 317
tandem cells, 431–433 photoluminescence intensity, 319
ultrafast forward electron injection in, GaAsN
423 in QW solar cells, 456, 457
GaAsP
edge-ray theorem, 512, 513 in QW solar cells, 12, 313, 455, 457,
EGing Photovoltaic Technology, 650 478–484
electronic conjugation, 343 GaInP/GaAs tandem cells, 309
September 17, 2014 11:9 Clean Electricity From Photovoltaics (2nd Edition) 9in x 6in b1809-index page 674
674 Index
Index 675
676 Index
Index 677