Fuel Cell Systems Explained
Fuel Cell Systems Explained
Third Edition
Andrew L. Dicks
Griffith University
Brisbane, Australia
David A. J. Rand
CSIRO Energy
Melbourne, Australia
This edition first published 2018
© 2018 John Wiley & Sons Ltd
Edition History
John Wiley & Sons Ltd (1e, 2000); John Wiley & Sons Ltd (2e, 2003)
All rights reserved. No part of this publication may be reproduced, stored in a retrieval system, or
transmitted, in any form or by any means, electronic, mechanical, photocopying, recording or otherwise,
except as permitted by law. Advice on how to obtain permission to reuse material from this title is available
at http://www.wiley.com/go/permissions.
The right of Andrew L. Dicks and David A. J. Rand to be identified as the authors of this work has been
asserted in accordance with law.
Registered Offices
John Wiley & Sons, Inc., 111 River Street, Hoboken, NJ 07030, USA
John Wiley & Sons Ltd, The Atrium, Southern Gate, Chichester, West Sussex, PO19 8SQ, UK
Editorial Office
The Atrium, Southern Gate, Chichester, West Sussex, PO19 8SQ, UK
For details of our global editorial offices, customer services, and more information about Wiley products
visit us at www.wiley.com.
Wiley also publishes its books in a variety of electronic formats and by print‐on‐demand. Some content that
appears in standard print versions of this book may not be available in other formats.
Limit of Liability/Disclaimer of Warranty
While the publisher and authors have used their best efforts in preparing this work, they make no
representations or warranties with respect to the accuracy or completeness of the contents of this work and
specifically disclaim all warranties, including without limitation any implied warranties of merchantability
or fitness for a particular purpose. No warranty may be created or extended by sales representatives, written
sales materials or promotional statements for this work. The fact that an organization, website, or product
is referred to in this work as a citation and/or potential source of further information does not mean that
the publisher and authors endorse the information or services the organization, website, or product may
provide or recommendations it may make. This work is sold with the understanding that the publisher is not
engaged in rendering professional services. The advice and strategies contained herein may not be suitable
for your situation. You should consult with a specialist where appropriate. Further, readers should be aware
that websites listed in this work may have changed or disappeared between when this work was written
and when it is read. Neither the publisher nor authors shall be liable for any loss of profit or any other
commercial damages, including but not limited to special, incidental, consequential, or other damages.
Library of Congress Cataloging‐in‐Publication Data
Names: Dicks, Andrew L., author. | Rand, David A. J., 1942– author.
Title: Fuel cell systems explained / Andrew L. Dicks, Griffith University, Brisbane, Australia,
David A. J. Rand, CSIRO Energy, Melbourne, Australia.
Description: Third edition. | Hoboken, NJ, USA : Wiley, [2018] | Includes bibliographical references
and index. |
Identifiers: LCCN 2017054489 (print) | LCCN 2017058097 (ebook) | ISBN 9781118706978 (pdf ) |
ISBN 9781118706961 (epub) | ISBN 9781118613528 (cloth)
Subjects: LCSH: Fuel cells.
Classification: LCC TK2931 (ebook) | LCC TK2931 .L37 2017 (print) | DDC 621.31/2429–dc23
LC record available at https://lccn.loc.gov/2017054489
Cover design by Wiley
Cover images: Top Image: © Iain Masterton/Alamy Stock Photo; Bottom Image: Courtesy of FuelCell
Energy, Inc.
Set in 10/12pt Warnock by SPi Global, Pondicherry, India
Printed in the UK by Bell & Bain Ltd, Glasgow
10
9
8
7
6
5
4
3
2
1
v
Contents
Brief Biographies xiii
Preface xv
Acknowledgments xvii
Acronyms and Initialisms xix
Symbols and Units xxv
1
1.1
1.2
1.3
1.4
1.5
1.6
1.7
1.7.1
1.7.2
1.7.3
1.8
1.9
1.10
2
2.1
2.2
2.3
2.4
2.5
2.5.1
2.5.2
2.5.3
2.5.4
2.6
Introducing Fuel Cells 1
Historical Perspective 1
Fuel‐Cell Basics 7
Electrode Reaction Rates 9
Stack Design 11
Gas Supply and Cooling 14
Principal Technologies 17
Mechanically Rechargeable Batteries and Other Fuel Cells 19
Metal–Air Cells 20
Redox Flow Cells 20
Biological Fuel Cells 23
Balance‐of‐Plant Components 23
Fuel‐Cell Systems: Key Parameters 24
Advantages and Applications 25
Further Reading 26
Efficiency and Open‐Circuit Voltage 27
Open‐Circuit Voltage: Hydrogen Fuel Cell 27
Open‐Circuit Voltage: Other Fuel Cells and Batteries 31
Efficiency and Its Limits 32
Efficiency and Voltage 35
Influence of Pressure and Gas Concentration 36
Nernst Equation 36
Hydrogen Partial Pressure 38
Fuel and Oxidant Utilization 39
System Pressure 39
Summary 40
Further Reading 41
vi
Contents
3
Operational Fuel‐Cell Voltages 43
3.1
3.2
3.3
3.4
3.4.1
3.4.2
3.4.3
3.5
3.6
3.7
3.8
3.9
3.10
3.10.1
3.10.2
3.10.3
Fundamental Voltage: Current Behaviour 43
Terminology 44
Fuel‐Cell Irreversibilities 46
Activation Losses 46
The Tafel Equation 46
The Constants in the Tafel Equation 48
Reducing the Activation Overpotential 51
Internal Currents and Fuel Crossover 52
Ohmic Losses 54
Mass‐Transport Losses 55
Combining the Irreversibilities 57
The Electrical Double-Layer 58
Techniques for Distinguishing Irreversibilities 60
Cyclic Voltammetry 60
AC Impedance Spectroscopy 61
Current Interruption 65
Further Reading 68
4
Proton‐Exchange Membrane Fuel Cells 69
4.1
4.2
4.2.1
4.2.2
4.2.3
4.2.4
4.2.5
4.3
4.3.1
4.3.2
4.3.2.1
4.3.2.2
4.3.2.3
4.3.2.4
4.3.2.5
4.3.2.6
4.3.3
4.3.4
4.3.5
4.4
4.4.1
4.4.2
4.4.3
4.4.4
4.4.5
4.4.6
4.5
Overview 69
Polymer Electrolyte: Principles of Operation 72
Perfluorinated Sulfonic Acid Membrane 72
Modified Perfluorinated Sulfonic Acid Membranes 76
Alternative Sulfonated and Non‐Sulfonated Membranes 77
Acid–Base Complexes and Ionic Liquids 79
High‐Temperature Proton Conductors 80
Electrodes and Electrode Structure 81
Catalyst Layers: Platinum‐Based Catalysts 82
Catalyst Layers: Alternative Catalysts for Oxygen Reduction 85
Macrocyclics 86
Chalcogenides 87
Conductive Polymers 87
Nitrides 87
Functionalized Carbons 87
Heteropolyacids 88
Catalyst Layer: Negative Electrode 88
Catalyst Durability 88
Gas‐Diffusion Layer 89
Water Management 92
Hydration and Water Movement 92
Air Flow and Water Evaporation 94
Air Humidity 96
Self‐Humidified Cells 98
External Humidification: Principles 100
External Humidification: Methods 102
Cooling and Air Supply 104
Contents
4.5.1
4.5.2
4.5.3
4.6
4.6.1
4.6.2
4.6.3
4.6.4
4.6.5
4.6.6
4.7
4.7.1
4.7.2
4.7.2.1
4.7.3
4.8
4.8.1
4.8.2
4.9
4.9.1
4.9.2
4.9.3
4.10
4.10.1
4.10.2
4.10.3
4.11
Cooling with Cathode Air Supply 104
Separate Reactant and Cooling Air 104
Water Cooling 105
Stack Construction Methods 107
Introduction 107
Carbon Bipolar Plates 107
Metal Bipolar Plates 109
Flow‐Field Patterns 110
Other Topologies 112
Mixed Reactant Cells 114
Operating Pressure 115
Technical Issues 115
Benefits of High Operating Pressures 117
Current 117
Other Factors 120
Fuel Types 120
Reformed Hydrocarbons 120
Alcohols and Other Liquid Fuels 121
Practical and Commercial Systems 122
Small‐Scale Systems 122
Medium‐Scale for Stationary Applications 123
Transport System Applications 125
System Design, Stack Lifetime and Related Issues 129
Membrane Degradation 129
Catalyst Degradation 129
System Control 129
Unitized Regenerative Fuel Cells 130
Further Reading 132
5
Alkaline Fuel Cells 135
Principles of Operation 135
System Designs 137
Circulating Electrolyte Solution 137
Static Electrolyte Solution 140
Dissolved Fuel 142
Anion‐Exchange Membrane Fuel Cells 144
Electrodes 147
Sintered Nickel Powder 147
Raney Metals 147
Rolled Carbon 148
Catalysts 150
Stack Designs 151
Monopolar and Bipolar 151
Other Stack Designs 152
Operating Pressure and Temperature 152
Opportunities and Challenges 155
Further Reading 156
5.1
5.2
5.2.1
5.2.2
5.2.3
5.2.4
5.3
5.3.1
5.3.2
5.3.3
5.3.4
5.4
5.4.1
5.4.2
5.5
5.6
vii
viii
Contents
6
Direct Liquid Fuel Cells 157
6.1
6.1.1
6.1.2
6.1.3
6.1.4
6.1.5
6.1.6
6.1.7
6.1.8
6.1.9
6.1.10
6.1.11
6.2
6.2.1
6.2.2
6.2.3
6.2.4
6.3
6.4
6.4.1
6.4.2
6.4.3
6.5
6.5.1
6.5.2
6.6
6.6.1
6.6.2
6.7
Direct Methanol Fuel Cells 157
Principles of Operation 160
Electrode Reactions with a Proton‐Exchange Membrane Electrolyte 160
Electrode Reactions with an Alkaline Electrolyte 162
Anode Catalysts 162
Cathode Catalysts 163
System Designs 164
Fuel Crossover 165
Mitigating Fuel Crossover: Standard Techniques 166
Mitigating Fuel Crossover: Prospective Techniques 167
Methanol Production 168
Methanol Safety and Storage 168
Direct Ethanol Fuel Cells 169
Principles of Operation 170
Ethanol Oxidation, Catalyst and Reaction Mechanism 170
Low‐Temperature Operation: Performance and Challenges 172
High‐Temperature Direct Ethanol Fuel Cells 173
Direct Propanol Fuel Cells 173
Direct Ethylene Glycol Fuel Cells 174
Principles of Operation 174
Ethylene Glycol: Anodic Oxidation 175
Cell Performance 176
Formic Acid Fuel Cells 176
Formic Acid: Anodic Oxidation 177
Cell Performance 177
Borohydride Fuel Cells 178
Anode Catalysts 180
Challenges 180
Application of Direct Liquid Fuel Cells 182
Further Reading 184
7
Phosphoric Acid Fuel Cells 187
7.1
7.2
7.2.1
7.2.2
7.2.3
7.2.3.1
7.2.3.2
7.2.3.3
7.3
7.3.1
7.3.2
7.3.3
7.3.4
7.4
High‐Temperature Fuel‐Cell Systems 187
System Design 188
Fuel Processing 188
Fuel Utilization 189
Heat‐Exchangers 192
Designs 193
Exergy Analysis 193
Pinch Analysis 194
Principles of Operation 196
Electrolyte 196
Electrodes and Catalysts 198
Stack Construction 199
Stack Cooling and Manifolding 200
Performance 201
Contents
7.4.1
7.4.2
7.4.3
7.4.4
7.5
Operating Pressure 202
Operating Temperature 202
Effects of Fuel and Oxidant Composition 203
Effects of Carbon Monoxide and Sulfur 204
Technological Developments 204
Further Reading 206
8
Molten Carbonate Fuel Cells 207
8.1
8.2
8.2.1
8.2.2
8.2.3
8.2.4
8.3
8.3.1
8.3.2
8.4
8.4.1
8.4.2
8.5
8.5.1
8.5.2
8.5.3
8.5.4
8.6
8.7
8.8
Principles of Operation 207
Cell Components 210
Electrolyte 211
Anode 213
Cathode 214
Non‐Porous Components 215
Stack Configuration and Sealing 215
Manifolding 216
Internal and External Reforming 218
Performance 220
Influence of Pressure 220
Influence of Temperature 222
Practical Systems 223
Fuel Cell Energy (USA) 223
Fuel Cell Energy Solutions (Europe) 225
Facilities in Japan 228
Facilities in South Korea 228
Future Research and Development 229
Hydrogen Production and Carbon Dioxide Separation 230
Direct Carbon Fuel Cell 231
Further Reading 234
9
Solid Oxide Fuel Cells 235
Principles of Operation 235
High‐Temperature (HT) Cells 235
Low‐Temperature (IT) Cells 237
Components 238
Zirconia Electrolyte for HT‐Cells 238
Electrolytes for IT‐Cells 240
Ceria 240
Perovskites 241
Other Materials 243
Anodes 243
Nickel‐YSZ 243
Cathode 245
Mixed Ionic–Electronic Conductor Anode
Cathode 247
Interconnect Material 247
Sealing Materials 248
9.1
9.1.1
9.1.2
9.2
9.2.1
9.2.2
9.2.2.1
9.2.2.2
9.2.2.3
9.2.3
9.2.3.1
9.2.3.2
9.2.3.3
9.2.4
9.2.5
9.2.6
246
ix
x
Contents
9.3
9.3.1
9.3.2
9.4
9.5
9.5.1
9.5.2
9.6
Practical Design and Stacking Arrangements 249
Tubular Design 249
Planar Design 251
Performance 253
Developmental and Commercial Systems 254
Tubular SOFCs 255
Planar SOFCs 256
Combined‐Cycle and Other Systems 258
Further Reading 260
10
Fuels for Fuel Cells 263
Introduction 263
Fossil Fuels 266
Petroleum 266
Petroleum from Tar Sands, Oil Shales and Gas Hydrates 268
Coal and Coal Gases 268
Natural Gas and Coal‐Bed Methane (Coal‐Seam Gas) 270
Biofuels 272
Basics of Fuel Processing 275
Fuel‐Cell Requirements 275
Desulfurization 275
Steam Reforming 277
Carbon Formation and Pre‐Reforming 280
Internal Reforming 281
Indirect Internal Reforming (IIR) 283
Direct Internal Reforming (DIR) 283
Direct Hydrocarbon Oxidation 284
Partial Oxidation and Autothermal Reforming 285
Solar–Thermal Reforming 286
Sorbent‐Enhanced Reforming 287
Hydrogen Generation by Pyrolysis or Thermal Cracking
of Hydrocarbons 289
Further Fuel Processing: Removal of Carbon Monoxide 290
Membrane Developments for Gas Separation 293
Non‐Porous Metal Membranes 293
Non‐Porous Ceramic Membranes 294
Porous Membranes 294
Oxygen Separation 295
Practical Fuel Processing: Stationary Applications 295
Industrial Steam Reforming 295
Fuel‐Cell Plants Operating with Steam Reforming of Natural Gas 296
Reformer and Partial Oxidation Designs 298
Conventional Packed‐Bed Catalytic Reactors 298
Compact Reformers 299
Plate Reformers and Microchannel Reformers 300
Membrane Reactors 301
Non‐Catalytic Partial Oxidation Reactors 302
10.1
10.2
10.2.1
10.2.2
10.2.3
10.2.4
10.3
10.4
10.4.1
10.4.2
10.4.3
10.4.4
10.4.5
10.4.5.1
10.4.5.2
10.4.6
10.4.7
10.4.8
10.4.9
10.4.10
10.4.11
10.5
10.5.1
10.5.2
10.5.3
10.5.4
10.6
10.6.1
10.6.2
10.6.3
10.6.3.1
10.6.3.2
10.6.3.3
10.6.3.4
10.6.3.5
Contents
10.6.3.6
10.7
10.8
10.8.1
10.8.2
10.8.3
10.8.4
10.9
10.9.1
10.9.2
10.10
10.10.1
10.10.2
10.10.3
10.10.4
Catalytic Partial Oxidation Reactors 303
Practical Fuel Processing: Mobile Applications 304
Electrolysers 305
Operation of Electrolysers 305
Applications 307
Electrolyser Efficiency 312
Photoelectrochemical Cells 312
Thermochemical Hydrogen Production and Chemical Looping 314
Thermochemical Cycles 314
Chemical Looping 317
Biological Production of Hydrogen 318
Introduction 318
Photosynthesis and Water Splitting 318
Biological Shift Reaction 320
Digestion Processes 320
Further Reading 321
11
Hydrogen Storage
11.1
11.2
11.3
11.3.1
11.3.2
11.3.3
11.3.4
11.4
11.5
11.6
11.6.1
11.6.2
11.6.3
11.7
11.7.1
11.7.2
11.8
11.9
323
Strategic Considerations 323
Safety 326
Compressed Hydrogen 327
Storage Cylinders 327
Storage Efficiency 329
Costs of Stored Hydrogen 330
Safety Aspects 330
Liquid Hydrogen 331
Reversible Metal Hydrides 333
Simple Hydrogen‐Bearing Chemicals 338
Organic Chemicals 338
Alkali Metal Hydrides 339
Ammonia, Amines and Ammonia Borane 340
Complex Chemical Hydrides 341
Alanates 342
Borohydrides 342
Nanostructured Materials 344
Evaluation of Hydrogen Storage Methods 347
Further Reading 350
12
The Complete System and Its Future
12.1
12.1.1
12.1.1.1
12.1.1.2
12.1.1.3
12.1.1.4
12.1.2
12.1.3
351
Mechanical Balance‐of‐Plant Components 351
Compressors 351
Efficiency 354
Power 356
Performance Charts 356
Selection 359
Turbines 361
Ejector Circulators 362
xi
xii
Contents
12.1.4
12.1.5
12.2
12.2.1
12.2.2
12.2.3
12.2.4
12.2.4.1
12.2.4.2
12.2.5
12.2.6
12.3
12.4
12.4.1
12.4.2
12.4.3
12.4.4
12.4.5
12.5
12.5.1
12.5.2
12.6
Fans and Blowers 363
Pumps 364
Power Electronics 365
DC Regulators (Converters) and Electronic Switches 366
Step‐Down Regulators 368
Step‐Up Regulators 370
Inverters 371
Single Phase 372
Three Phase 376
Fuel‐Cell Interface and Grid Connection Issues 378
Power Factor and Power Factor Correction 378
Hybrid Fuel‐Cell + Battery Systems 380
Analysis of Fuel‐Cell Systems 384
Well‐to‐Wheels Analysis 385
Power‐Train Analysis 387
Life‐Cycle Assessment 388
Process Modelling 389
Further Modelling 392
Commercial Reality 394
Back to Basics 394
Commercial Progress 395
Future Prospects: The Crystal Ball Remains Cloudy 397
Further Reading 399
Appendix 1
A1.1
A1.2
Appendix 2
A2.1
A2.2
A2.3
A2.4
A2.5
A2.6
Useful Fuel‐Cell Equations
405
Introduction 405
Oxygen and Air Usage 406
Exit Air Flow Rate 407
Hydrogen Usage 407
Rate of Water Production 408
Heat Production 409
Appendix 3
A3.1
A3.2
Calculations of the Change in Molar Gibbs Free Energy 401
Hydrogen Fuel Cell 401
Carbon Monoxide Fuel Cell 403
Calculation of Power Required by Air Compressor and Power Recoverable
by Turbine in Fuel‐Cell Exhaust 411
Power Required by Air Compressor 411
Power Recoverable from Fuel‐Cell Exhaust with a Turbine
Glossary of Terms 415
Index 437
412
xiii
Brief Biographies
Andrew L. Dicks
Andrew L. Dicks, PhD, CChem, FRSC,
was educated in England and graduated
from Loughborough University before
starting a career in the corporate laboratories of the UK gas industry. His first
research projects focused on heterogeneous catalysts in gas‐making processes,
for which he was awarded a doctorate in
1981. In the mid‐1980s, BG appointed
Andrew to lead a research effort on fuel
cells that was directed predominantly
towards molten carbonate and solid
oxide systems. The team pioneered the
application of process modelling to
fuel‐cell systems, especially those that
featured internal reforming. This work,
which was supported by the European
Commission during the 1990s, involved
collaboration with leading fuel‐cell
developers throughout Europe and
North America. In 1994, Andrew was
jointly awarded the Sir Henry Jones
(London) Medal of the Institution of Gas Engineers and Managers for his studies on
high‐temperature systems. He also took an interest in proton‐exchange membrane fuel
cells and became the chair of a project at the University of Victoria, British Columbia,
in which Ballard Power Systems was the industrial partner. In 2001, he was awarded a
Senior Research Fellowship at the University of Queensland, Australia, that enabled
further pursuit of his interest in catalysis and the application of nanomaterials in fuel‐
cell systems. Since moving to Australia, he has continued to promote hydrogen and
fuel‐cell technology, as director of the CSIRO National Hydrogen Materials Alliance
and as a director of the Australian Institute of Energy. He is now consulted on energy
and clean technology issues by governments and funding agencies worldwide.
xiv
Brief Biographies
David A. J. Rand
David A. J. Rand, AM, BA, MA,
PhD, ScD, FTSE, was educated
at the University of Cambridge
where, after graduation, he
conducted research on low‐
temperature fuel cells. In 1969,
he joined the Australian government’s CSIRO laboratories
in Melbourne. After further
exploration of fuel‐cell mechanisms and then electrochemical
studies of mineral beneficiation, he formed the CSIRO
Novel Battery Technologies
Group in the late 1970s and
remained its leader until 2003. He was one of the six scientists who established the US‐
based Advanced Lead–Acid Battery Consortium in 1992 and served as its manager in
1994. He is the co‐inventor of the UltraBatteryTM, which finds service in hybrid electric
vehicle and renewable energy storage applications. As a chief research scientist, he fulfilled the role of CSIRO’s scientific advisor on hydrogen and renewable energy until his
retirement in 2008. He remains active within the organisation as an Honorary Research
Fellow and has served as the chief energy scientist of the World Solar Challenge since its
inception in 1987. He was awarded the Faraday Medal by the Royal Society of Chemistry
(United Kingdom) in 1991, the UNESCO Gaston Planté Medal by the Bulgarian
Academy of Sciences in 1996 and the R.H. Stokes Medal by the Royal Australian
Chemical Institute in 2006. He was elected a fellow of the Australian Academy of
Technological Sciences and Engineering in 1998 and became a member of the Order
of Australia in 2013 for service to science and technological development in the field of
energy storage.
xv
Preface
Since publication of the first edition of Fuel Cell Systems Explained, three compelling
drivers have supported the continuing development of fuel‐cell technology, namely:
●
●
●
The need to maintain energy security in an energy‐hungry world.
The desire to reduce urban air pollution from vehicles.
The mitigation of climate change by lowering anthropogenic emissions of carbon
dioxide.
New materials for fuel cells, together with improvements in the performance and
lifetimes of stacks, are underpinning the emergence of the first truly commercial
systems in applications that range from forklift trucks to power sources for mobile phone
towers. Leading vehicle manufacturers have embraced the use of electric drivetrains
and now see hydrogen fuel cells complementing the new battery technologies that have
also emerged over the past few years. After many decades of laboratory development, a
global — but fragile — fuel‐cell industry is bringing the first products to market.
To assist those who are unfamiliar with fuel‐cell electrochemistry, Chapter 1 of this
third edition has been expanded to include a more detailed account of the evolution of
the fuel cell and its accompanying terminology. In the following chapters, extensive
revision of the preceding publication has removed material that is no longer relevant to
the understanding of modern fuel‐cell systems and has also introduced the latest
research findings and technological advances. For example, there are now sections
devoted to fuel‐cell characterization, new materials for low‐temperature hydrogen
and liquid‐fuelled systems, and a review of system commercialization. Separate
chapters on fuel processing and hydrogen storage have been introduced to emphasize
how hydrogen may gain importance both in future transport systems and in providing
the means for storing renewable energy.
The objective of each chapter is to encourage the reader to explore the subject in
more depth. For this reason, references have been included as footnotes when it is
necessary to substantiate or reinforce the text. To stimulate further interest, however,
some recommended further reading may be given at the end of a chapter.
There are now several books and electronic resources available to engineers and
scientists new to fuel‐cell systems. The third edition of Fuel Cell Systems Explained
does not intend to compete with specialist texts that can easily be accessed via the
Internet. Rather, it is expected that the book will continue to provide an introduction
and overview for students and teachers at universities and technical schools and act as
xvi
Preface
a primer for postgraduate researchers who have chosen to enter this field of technology.
Indeed, it is hoped that all readers — be they practitioners, researchers and students
in electrical, power, chemical and automotive engineering disciplines — will continue
to benefit from this essential guide to the principles, design and implementation of
fuel‐cell systems.
December 2017
Andrew L. Dicks, Brisbane, Australia
David A. J. Rand, Melbourne, Australia
xvii
Acknowledgments
As emphasized throughout this publication, the research and development of fuel cells
is highly interdisciplinary in that it encompasses many aspects of science and engineering. This fact is reflected in the number and diversity of companies and organizations
that have willingly provided advice and information or given permission to use their
images in the third edition of Fuel Cell Systems Explained. Accordingly, the authors are
indebted to the following contributors:
Avantica plc (formerly BG Technology Ltd), UK
Ballard Power Systems Inc., USA
CNR ITAE, Italy
Coregas, Australia
Cygnus Atratus, UK
Daimler AG, Germany
Doosan Fuel Cell, USA
Eaton Corporation, USA
Forschungszentrum Jülich GmbH, Germany
Fuel Cell Energy, USA
Horizon Fuel Cells, Singapore
Hydrogenics Corporation, Canada
Hyundai Motor Company, Australia Pty Ltd
Intelligent Energy, UK
International Fuel Cells, USA
ITM Power, UK
Johnsons Matthey plc, UK
Kawasaki Heavy Industries, Japan
Kyocera, Japan
NDC Power, USA
Osaka Gas, Japan
Proton Energy Systems, USA
Proton Motor Systems, GmbH, Germany
Redflow Ltd, Australia
Serenergy, Denmark
Siemens Westinghouse Power Corporation, USA
In addition, the authors acknowledge the work of James Larminie, who instigated the
first edition of this book, as well as the assistance of others engaged in the advancement
xviii
Acknowledgments
of fuel cells, namely, John Appleby (Texas A&M University, USA), Nigel Brandon and
David Hart (Imperial College, UK), John Andrews (RMIT University, Australia), Evan
Gray (Griffith University, Australia), Ian Gregg (Consultant, Australia) and Chris
Hodrien (University of Warwick, UK).
The authors also wish to express their thanks for the support and encouragement
given by family, friends and colleagues during the course of this project.
xix
Acronyms and Initialisms
ABPBI
AC
ADP
AEM
AEMFC
AES
AFC
AMFC
ANL
APEMFC
APU
ASR
phosphoric acid doped poly(2,5‐benzimidazole)
alternating current
adenosine 5’-triphosphate
alkaline‐electrolyte membrane
alkaline‐electrolyte membrane fuel cell
air‐electrode supported
alkaline fuel cell
anion‐exchange membrane fuel cell
Argonne National Laboratory
alkaline proton‐exchange membrane fuel cell
auxiliary power unit
area specific resistance
BCN
BG
BIMEVOX
BOP
BPS
BSF
Dutch Fuel Cell Corporation
British Gas
bismuth metal vanadium oxide (Bi4V2O11)
balance-of-plant
Ballard Power Systems
Boudouard Safety Factor
CAN bus
CBM
CCS
CFCL
CGO
CHP
CLC
CNR
CNT
CODH-1
CPE
CPO
CRG
CSG
CSIRO
Controller Area Network
coal‐bed methane
carbon capture and storage
Ceramic Fuel Cells Ltd
cerium–gadolinium oxide (same as GDC)
combined heat and power
chemical looping combustion
Consiglio Nazionale delle Ricerche (Italy)
carbon nanotube
carbon monoxide dehydrogenase
constant phase element
catalytic partial oxidation
catalytic rich gas
coal‐seam gas
Commonwealth Scientific and Industrial Research Organisation
xx
Acronyms and Initialisms
CSO
CSZ
CV
CVD
cerium‐samarium oxide (same as SDC)
calcia‐stabilized zirconia
cyclic voltammetry
chemical vapour deposition
DBFC
DC
DCFC
DEFC
DEGFC
DFAFC
DFT
DG
DIR
DIVRR
DLFC
DMFC
DOE
DPFC
DPFC(2)
DSSC
direct borohydride fuel cell
direct current
direct carbon fuel cell
direct ethanol fuel cell
direct ethylene glycol fuel cell
direct formic acid fuel cell (also formic acid fuel cell, FAFC)
density functional theory
distributed generator
direct internal reforming
directly irradiated, volumetric receiver–reactor
direct liquid fuel cell
direct methanol fuel cell
Department of Energy (United States)
direct propanol fuel cell
direct propan‐2‐ol fuel cell
dye‐sensitized solar cell
EC
ECN
EFOY
EIS
EPFL
EU
EVD
EW
evaporatively cooled
Energy Research Centre of the Netherlands
Energy for You
electrochemical impedance spectroscopy
Swiss Federal Institute of Technology
European Union
electrochemical vapour deposition
membrane equivalent weight
FCE
FCES
FCV
FRA
FT
Fuel Cell Energy Inc.
Fuel Cell Energy Solutions GmbH
fuel cell vehicle
frequency response analyser
Fischer–Tropsch
GDC
GDL
GE
GHG
GM
GPS
GTL
GTO
gadolinium‐doped ceria/gadolinia‐doped ceria (same as CGO)
gas-diffusion layer
General Electric
greenhouse gas
General Motors
Global Positioning System
gas‐to‐liquid
gate turn‐off (thyristor)
HAZID
HAZOP
hazard identification
hazard and operability study
Acronyms and Initialisms
HCNG
HDS
HEMFC
HEV
HHV
HOR
HPE
hydrogen-compressed natural gas
hydrodesulfurization
hydroxide‐exchange polymer membrane fuel cell
hybrid electric vehicle
higher heating value
hydrogen oxidation reaction
high‐pressure proton‐exchange membrane electrolyser
IBFC
ICE
ICEV
IFC
IGBT
IHI
IHP
IIR
ITM
IT‐SOFC
IUPAC
indirect borohydride fuel cell
internal combustion engine
internal combustion engine vehicle
International Fuel Cells
insulated‐gate bipolar transistor
Ishikawajima‐Harima Heavy Industries Co., Ltd
inner Helmholtz plane
indirect internal reforming (also known as ‘integrated reforming’)
ion transport membrane, also refers to company ITM Power
intermediate‐temperature solid oxide fuel cell
International Union of Pure and Applied Chemistry
KEPCO
KIST
Korea Electric Power Corporation
Korea Institute of Science and Technology
LAMOX
LCA
lanthanum molybdate (La2Mo2O9)
life‐cycle assessment (also known as ‘life‐cycle analysis’ and ‘cradle‐to‐grave
analysis’)
LCOE
levelized cost of electricity
LH2
liquid hydrogen
LHV
lower heating value
LNG
liquefied natural gas
LPG
liquefied petroleum gas
LSCF
lanthanum strontium cobaltite ferrite
LSCV
strontium‐doped lanthanum vanadate
LSGM
lanthanum gallate (LaSrGaMgO3)
LSM
strontium‐doped lanthanum manganite
LT‐SOFC low‐temperature solid oxide fuel cell
MCFC
MCR
MEA
MEMS
METI
MFC
MFF
MHPS
MIEC
MOF
MOSFET
molten carbonate fuel cell
microchannel reactor
membrane–electrode assembly
microelectromechanical systems
Ministry of Economy, Trade and Industry (Japan)
microbial fuel cell
mass flow factor
Mitsubishi Hitachi Power Systems
mixed ionic–electronic conductor (oxides)
metal–organic framework
metal‐oxide‐semiconductor field‐effect transistor
xxi
xxii
Acronyms and Initialisms
MPMDMS
MRFC
MSW
MTBF
MWCNT
(3‐mercaptopropyl)methyldimethoxysilane
mixed‐reactant fuel cell
municipal solid waste
mean time between failures
multiwalled carbon nanotube
NADP
NASA
NCPO
NEDO
NOMO
NTP
nicotinamide adenine dinucleotide phosphate
National Aeronautics and Space Administration
non-catalytic partial oxidation
New Energy Development Organization (Japan)
Notice of Market Opportunities
normal temperature and pressure
OCV
OEM
OER
OHP
ORR
open‐circuit voltage
original equipment manufacturer
oxygen evolution reaction
outer Helmholtz plane
oxygen reduction reaction
P2G
P3MT
PAFC
PANI
PAR
PBI
PBSS
PC
PCT
PEC
PEMFC
PET
PF
PFD
PFSA
plc
POX
PPA
PPBP
Ppy
PROX
PrOx
PSA
PTFE
PV
PWM
power‐to‐gas
poly(3‐methylthiophene)
phosphoric acid fuel cell
polyaniline
photosynthetically active radiation
polybenzimidazole
poly(benzylsulfonic acid)siloxane
phthalocyanine
pressure composition isotherm
photoelectrochemical cell
proton‐exchange membrane fuel cell (also called ‘polymer electrolyte
membrane fuel cell’ and same as SPEFC and SPFC)
polyethylene terephthalate
power factor, also PFC power factor correction
process flow diagram
perfluorinated sulfonic acid
programmable logic controller
partial oxidation
polyphosphoric acid
poly(1,4‐phenylene), poly(4 phenoxybenzoyl‐1,4‐phenylene)
polypyrrole
preferential oxidation
preferential oxidation reactor
pressure swing adsorption
polytetrafluoroethylene
photovoltaic
pulse width modulation
QA
quaternary ammonium
Acronyms and Initialisms
RDE
RFB
RH
RHE
RRDE
RSF
rotating disc electrode
redox flow battery
relative humidity
reversible hydrogen electrode
rotating ring‐disc electrode
rotational speed factor
SATP
SCG
SCT‐CPO
SDC
SECA
SFCM
SHE
SI
SLM
SMR
SNG
SOFC
m-SPAEEN-60
SPEEK
SPEFC
SPFC
SPOF
STP
SWPC
standard ambient temperature and pressure
simulated coal gas
short contact time catalytic partial oxidation
samarium‐doped ceria/samaria‐doped ceria (same as CSO)
Solid State Energy Conversion Alliance
standard cubic foot per minute
standard hydrogen electrode
International System of Units (French: Système international d’unités)
standard litre per minute
steam reforming reaction
substitute natural gas (also synthetic natural gas)
solid oxide fuel cell
sulfonated poly(arylene ether ether nitrile)
sulfonated polyether ether ketone
solid polymer electrolyte fuel cell (same as PEMFC)
solid polymer fuel cell (same as PEMFC)
single point of failure
standard temperature and pressure
Siemens Westinghouse Power Corporation
TAA
THT
TMPP
TPP
TPTZ
TTW
tetraazaannulene
tetrahydrothiophene
tetramethoxyphenylporphyrin
tetraphenylporphyrin
2, 4, 6‐tris(2‐pyridyl)‐1,3,5‐triazine
tank‐to‐wheel
UCC
UK
ULP
UPS
URFC
USA
USB
UTC
UV
Union Carbide Corporation
United Kingdom
unleaded petrol
uninterruptible power system; also uninterruptible power supply
unitized regenerative fuel cell
United States of America
universal serial bus
United Technologies Corporation
ultraviolet
WGS
WTT
WTW
water–gas shift
well‐to‐tank
well‐to‐wheels
XPS
X‐ray photoelectron spectroscopy
xxiii
xxv
Symbols and Units
Subunits
d
c
m
μ
n
A
A
Multiple units
deci
centi
milli
micro
nano
10−1
10−2
10−3
10−6
10−9
k
M
G
T
P
kilo
mega
giga
tera
peta
103
106
109
1012
1015
ampere
electrode area (cm2), also coefficient in natural logarithm form of the Tafel
equation
Ah ampere hour
a
chemical activity; also coefficient in base 10 logarithm form of the Tafel
equation
ax
chemical activity of species x
atm atmosphere (=101.325 kPa)
B
exergy (J)
ΔB change in exergy (J)
bbl barrel of oil: 35 imperial gallons (159.113 L), or 42 US gallons (158.987 L)
bar unit of pressure (=100 kPa)
bhp brake horsepower (=745.7 W)
C
constant in various equations; also coulomb (=1A s), the unit of electric charge
°C
degree Celsius
CP specific heat capacity at constant pressure (J kg−1 K−1)
CV specific heat capacity at constant volume (J kg−1 K−1)
cP
molar heat capacity at constant pressure (J mol−1 K−1)
cV
molar heat capacity at constant volume (J mol−1 K−1)
cm centimetre
Dm diffusion coefficient (m2 s−1)
d
separation of charge layers in a capacitor (mm)
E
electrode potential (V)
E°
standard electrode potential (V)
Er
reversible electrode potential (V)
E r standard reversible electrode potential (V)
xxvi
Symbols and Units
EW
e−
ΔEact
F
F
ft
G
ΔG
ΔG°
G f
G f
g
g
g
gf
g f
g
g
H
ΔH
ΔH°
H f
H f
h
h
h
hf
h f
h
IR e/
IR t/
I
i
ic
il
io
J
K
L
MFF
m
ṁ
mx
mEq
mol
N
0003367618.INDD 26
(membrane) equivalent weight
electron, or the charge on one electron (=1.602 × 10−19 coulombs)
activation overpotential (V)
farad, unit of electrical capacitance (s4 A2 m−2 kg−1)
Faraday constant (=96 458 coulombs mol−1)
foot (linear measurement = 305 mm)
Gibbs free energy (J)
change in Gibbs free energy (J)
change in standard Gibbs free energy (J)
standard Gibbs free energy of formation (J)
change in standard Gibbs free energy of formation (J)
molar Gibbs free energy (J mol−1)
change in molar Gibbs free energy (J mol−1)
change in standard molar Gibbs free energy (J mol−1)
change in molar Gibbs free energy of formation (J mol−1)
change in standard molar Gibbs free energy of formation (J mol−1)
gram
acceleration due to gravity (m s−2)
enthalpy (J)
change in enthalpy (J)
change in standard enthalpy (J)
standard enthalpy of formation (J)
change in standard enthalpy (heat) of formation (J)
molar enthalpy (J mol−1)
change in molar enthalpy (J mol−1)
change in standard molar enthalpy (J mol−1)
change molar enthalpy of formation (J mol−1)
change in standard molar enthalpy of formation (J mol−1)
hour
resistive loss in electrolyte (Ω)
total resistive loss in electrodes (Ω)
current (A)
current density, i.e., current per unit area (usually expressed in mA cm−2)
crossover current (A)
limiting current density (usually expressed in mA cm−2)
exchange-current density (usually expressed in mA cm−2)
joule (=1 W s)
kelvin (used as a measure of absolute temperature)
litre
mass flow factor (kg s−1 K1/2 bar−1)
metre
mass flow rate, e.g., of gas (kg s−1) or of a liquid (ml min−1)
mass of substance x (g)
milliequivalent (weight) (mg L−1)
mole, i.e., mass of 6.022 × 1023 elementary units (atoms, molecules, etc.) of a
substance
newton (unit of force = 1 kg m s−2)
2/24/2018 9:01:39 AM
Symbols and Units
N
NA
N‐m3
n
ni
n x
P
Pe
P°
PSAT
Px
Pa
ppb
pH
ppm
R
R/
RDS,on
RH
®
r
S
S
ΔS
ΔS°
S f
∆S f
s
∆s
s
∆s f
∆ s f
s
SLM
T
TM
t
t1/2
V
Vc
Vr
Vr
ΔVgain
ΔVloss
vol.%
rotor speed of fan (revolutions per minute)
Avogadro’s number, 6.022140857 × 1023
normal cubic metre of gas (i.e., that measured at NTP)
number of units (electrons, atoms, molecules) involved in a chemical or electrochemical reaction; also number of cells in fuel‐cell stack
number of units or moles of species i
molar flow rate of species x (mol s−1)
pressure (in Pa, or bar)
power (W), only used when context is clear that pressure is not under discussion
standard pressure (=100 kPa)
saturated vapour pressure
partial pressure of species x
pascal (1 Pa = 1 N m−2 = 9.869 × 10−6 atm)
parts per billion
numerical scale used to specify the acidity or basicity of an aqueous solution
parts per million
gas constant (=8.1345 J K−1 mol−1)
resistance (Ω)
internal resistance of a transistor
relative humidity (%); also denoted by the symbol ϕ (v.i.)
registered trademark/copyright
area specific resistance (Ω cm2)
siemens, unit of conductance (Ω−1)
entropy (J K−1)
change in entropy (J K−1)
change in standard entropy (J K−1)
standard entropy of formation (J K−1)
change in standard entropy of formation (J K−1)
molar entropy (J K−1 mol−1)
change in molar entropy (J K−1 mol−1)
change in standard molar entropy (J mol−1)
change in molar entropy of formation (J mol−1)
change in standard molar entropy of formation (J mol−1)
second
standard litre per minute
temperature
trademark
tonne
half‐life
volt
cell voltage (V)
reversible cell voltage; also known as ‘open‐circuit voltage’ (V)
reversible cell voltage (V) under standard conditions of temperature (298.15 K)
and pressure (101.325 kPa)
voltage gain (V)
voltage loss (V)
volume percent
xxvii
xxviii
Symbols and Units
W
W′
W
Wel
Wth
Wh
wt.%
xi
Z
z
work done, e.g., in compressing a gas (J)
isentropic work (J)
watt
watt, electrical power
watt, thermal power
watt‐hour
weight percent
mole fraction of species i in solution
impedance (Ω)
number of units (electrons, atoms, molecules) involved in a chemical or electrochemical reaction
α
γ
δm
ɛ
ξ
η
η+
η−
ηC
ηf
ϑ
λ
μf
μi
μ
ϕ
ρ
ω
charge transfer coefficient
ratio of the specific heats of a gas CP:CV
thickness of proton exchange membrane (cm)
electrical permittivity (F m−1)
electro‐osmotic coefficient
electrode overpotential (V); also efficiency (%) (e.g., of a fuel cell)
overpotential at a positive electrode (V)
overpotential at a negative electrode (V)
isentropic compressor efficiency (%)
fuel utilization coefficient (%), a ‘figure of merit’ for DMFCs
phase angle
stoichiometric ratio
fuel utilization coefficient
chemical potential of species i (J kg−1 or J mol−1)
gas viscosity (centipoise, cP = 0.001 kg m−1 s−1)
relative humidity (usually expressed as a percentage); also denoted by RH
gas density (kg m−3)
humidity ratio, also known as ‘absolute humidity’ and ‘specific humidity’; symbol also used for radial frequency
ohm
Ω
1
1
Introducing Fuel Cells
1.1
Historical Perspective
This book is an introduction to fuel‐cell systems; it aims to provide an understanding
of the technology — what it is, how it works and what are its applications. Essentially,
a fuel cell can be defined as a device that produces electrical power directly from a fuel
via an electrochemical process. In some respects, this operation is similar to that of a
conventional battery except that the reactants are stored outside the cell. Therefore,
the performance of the device is limited only by the availability of the fuel and oxidant
supply and not by the cell design. For this reason, fuel cells are rated by their power
output (kW) rather than by their capacity (kWh).
Before addressing the technology in depth, it is necessary to understand that by virtue
of being electrochemical, fuel cells have both chemical and electrical characteristics.
Accordingly, their development has been inextricably linked with the development of
electrochemistry as a distinct branch of physical chemistry.
At the beginning of the 19th century, it was recognized that an ‘electrochemical cell’
(nowadays, commonly called a ‘battery’) could be made by placing two dissimilar metals
in an aqueous salt solution. This discovery was made by Alessandro Volta, the professor
of experimental physics at Pavia University, who constructed a pile of alternating discs
of copper (or silver or brass) and zinc (or tin) that were separated by pasteboard discs
(or ‘any other spongy matter’) soaked in brine. When the top and bottom of the pile
were connected by a wire, the assembly delivered, for the first time in history, a more or
less steady flow of electricity. Volta introduced the terms ‘electric current’ and ‘electromotive force’, the latter to denote the physical phenomenon that causes the current to
flow. In due course, he conveyed his findings in a letter dated 20 March 1800 to Joseph
Banks, the then president of the Royal Society. Known as the ‘Volta (or Voltaic) pile’, this
was the first ‘primary’ (or non‐rechargeable) power source, as opposed to a ‘secondary’
(or rechargeable) power source.
Sir Humphry Davy, who was working at the Royal Institution in London, soon recognized that the Volta pile produces electricity via chemical reactions at the metal–
solution interfaces — hydrogen is evolved on the ‘positive’ copper disc, and zinc is
consumed at the ‘negative’ disc. Indeed, this recognition of the relationship between
chemical and electrical effects prompted Davy to coin the word ‘electrochemical’, from
which sprang the science of ‘electrochemistry’. He gave warning that Volta’s work was
‘an alarm bell to experimenters all over Europe’. His prediction was soon to be verified.
Fuel Cell Systems Explained, Third Edition. Andrew L. Dicks and David A. J. Rand.
© 2018 John Wiley & Sons Ltd. Published 2018 by John Wiley & Sons Ltd.
2
Fuel Cell Systems Explained
Volta had sent his letter to the Royal Society in two parts because he anticipated
problems with its delivery given that correspondence from Italy had to pass through
France, which was then at war with Britain. While waiting for the second part to
arrive, Joseph Banks had shown the first few pages to Anthony Carlisle (a fashionable
London surgeon) who, in turn, with the assistance William Nicholson (a competent
amateur scientist) assembled on 30 April 1800 the first pile to be constructed in
England. Almost immediately, on 2 May 1800, the two investigators found that the
current from their device when passed through a dilute salt solution via two platinum
wires was capable of decomposing water into its constituents — hydrogen at one wire
and oxygen at the other. Details of the discovery were published in Nicholson’s own
journal in July of the same year. Thus, the new technique of ‘molecular splitting’ — to
be coined ‘electrolysis’ by Michael Faraday much later in 1834 and derived from the
Greek ‘lysis’ = separation — was demonstrated before Volta’s own account of the pile
was made public in September 1800. A schematic representation of the process is
shown in Figure 1.1a.
It was left to Michael Faraday, Davy’s brilliant student, to identify the mechanisms of
the processes that take place within ‘electrolytic’ cells and to give them a quantitative
basis. In addition, he was also the guiding force behind the nomenclature that is still in
use today. First, Faraday with the assistance of Whitlock Nicholl (his personal physician
and accomplished linguist) devised the name ‘electrode’ to describe a solid substance at
which an electrochemical reaction occurs and ‘electrolyte’ to describe the chemical
compound that provides an electrically conductive medium between electrodes. (Note
that in the case of dissolved materials, it is fundamentally incorrect to refer to the
‘electrolyte solution’ as the ‘electrolyte’; nevertheless, the latter terminology has become
common practice.) To distinguish between the electrode by which conventional current
(i.e., the reverse flow of electrons) enters an electrolytic cell and the electrode by which
(a) Electrolysis cell
Current
(b) Fuel cell
–
Current
+
External
power source
e–
e–
+
–
–
External
load
e–
+
–
–
Anion
Anion
+
+
Cation
Cathode
e–
Cation
Anode
Electrolyte solution
Anode
Cathode
Electrolyte solution
Figure 1.1 Terminology employed in operation of (a) electrolysis cells and (b) fuel cells.
Introducing Fuel Cells
it leaves, Faraday sought the assistance of the polymath William Whewell, the Master of
Trinity College at the University of Cambridge. In a letter dated 24 April 1834, he asked
Whewell:
‘Can you help me out to two good names not depending upon the idea of a current
in one direction only or upon positive or negative?’
In other words, he wanted terms that would be unaffected by any later change in the
convention adopted for the direction of current. Eventually, they settled on calling the
positive electrode an ‘anode’ and the negative electrode a ‘cathode’, which were coined
from Greek ‘ano‐dos’ (‘upwards’–‘a way’) to represent the path of electrons from the
positive electrode to the negative and ‘katho‐dos’ (‘downwards’–‘a way’) to represent
the counter direction. For an electrolytic cell, then, the anode is where the current
enters the electrolyte and the cathode is where the current leaves the electrolyte. Thus
the positive electrode sustains an oxidation (or ‘anodic’) reaction with the liberation of
electrons, while a reduction (or ‘cathodic’) reaction takes place at the negative electrode
with the uptake of electrons.
With use of the Greek neutral present participle ‘ion’ — ‘a moving thing’ — to describe
the migrating particles in electrolysis, two further terms were obtained, namely, ‘anion’,
i.e., the negatively charged species that goes to the anode against the current (or with
the flow of negative charge), and ‘cation’, i.e., the positively charged species that goes to
the cathode with the current (or against the flow of negative charge). The operation of
an electrolysis cell is shown in Figure 1.1a. It should be noted that the anode–cathode
terminology for an ‘electrolytic cell’ applies to a ‘battery under charge’ (secondary
system).
A fuel cell operates in the reverse manner to an electrolysis cell, i.e., it is a ‘galvanic’
cell that spontaneously produces a voltage (similar to a ‘battery under discharge’). The
anode of the electrolysis cell now becomes the cathode and the cathode becomes the
anode; see Figure 1.1b. Nevertheless, the directions of the migration of anions and cations with respect to current flow are unchanged such that the positive electrode remains
a positive electrode and the negative electrode remains a negative electrode. Thus, in a
fuel cell, the fuel is always oxidized at the anode (positive electrode), and the oxidant is
reduced at the cathode (negative electrode).
There is some debate over who discovered the principle of the fuel cell. In a letter
written in December 1838 and published on page 43 of the January issue of the January–
June 1839 Volume XIV of The London and Edinburgh Philosophical Magazine and
Journal of Science, the German scientist Christian Friedrich Schönbein described his
investigations on fluids that were separated from each other by a membrane and
connected to a galvanometer by means of platina wires. In the 10th of 14 reported tests,
one compartment contained dilute sulfuric acid that held some hydrogen, whereas the
other compartment contained dilute sulfuric acid that was exposed to air. Schönbein
detected a current and concluded that this was caused ‘by the combination of hydrogen
with (the) oxygen (contained dissolved in water)’. This discovery was largely overlooked,
however, after the publication of a letter from William Robert Grove, a Welsh lawyer
and a scientist at the Royal Institution; see Figure 1.2a. The letter, which was dated
14 December 1838, appeared on page 127 of the February issue of the aforementioned
Volume XIV and described his evaluation of electrode and electrolyte materials for use
3
4
Fuel Cell Systems Explained
(a)
(b)
Ox
Hy
Current
Ox
Hy
Ox
Hy
Ox
Hy
Ox
Hy
Figure 1.2 (a) William Robert Grove (1811–1896) and (b) Grove’s sketch of four cells of his gaseous
voltaic battery’ (1842). (Source: https://commons.wikimedia.org/w/index.php?curid=20390734.Used
under CC BY‐SA 3.0; https://creativecommons.org/licenses/by‐sa/3.0/.)
in batteries. Unfortunately, the order in which these two letters had been written is
unknown as Schönbein did not date his letter in full — he gave the month, but not the
day. In fact, this chronology is of little importance given the following postscript that
Grove had added to his letter in January 1839:
‘I should have pursued these experiments further, and with other metals, but was
led aside by some experiments with different solutions separated by a diaphragm
and connected by platinum plates; in many of these I have been anticipated.’
In the same postscript, Grove went on to speculate that by connecting such cells in
series sufficient voltage could be created to decompose water (by electrolysis).
Grove carried out many experiments that demonstrated the principle of the fuel cell.
In 1842, he realized that the reaction at the electrodes was dependent on an area of
contact between the gas reactant and a layer of liquid that was sufficiently thin to allow
the gas to diffuse to the solid electrode (today, this requirement is commonly related to
the formation of a ‘three‐phase boundary’ or ‘triple‐point junction’ where gas, electrolyte
and electrocatalyst come into simultaneous contact, v.i.). At that time, Grove was the
professor of experimental chemistry at the London Institution in Finsbury Circus,
and in the same communication he reported the invention of a ‘gaseous voltaic battery’.
The device employed two platinized platinum electrodes (to increase the real surface
area), and a series of fifty such pairs when semi‐immersed in dilute sulfuric acid solution was found ‘to whirl round’ the needle of a galvanometer, to give a painful shock to
five persons joining hands, to give a brilliant spark between charcoal points, and to
decompose hydrochloric acid, potassium iodide and acidulated water. An original
sketch of four such cells is reproduced in Figure 1.2b. It was also found that 26 cells
were the minimum number required to electrolyse water. Grove had indeed realized
Introducing Fuel Cells
the desire expressed in his 1839 postscript in that he had achieved the beautiful
symmetry inherent in the ‘decomposition of water by means of its composition’.
The aforementioned apparatus became widely recognized as the first fuel cell and
Grove was designated as the ‘Father of the Fuel Cell’. Historically, this title is not fully
justified. More accurately, Schönbein should be credited with the discovery of the fuel‐
cell effect in 1838 and Grove with the invention of the first working prototype in 1842.
Happily, such accreditations were of little concern to the two scientists and they became
close friends. For almost 30 years, they exchanged ideas and developments via a dynamic
correspondence and visited each other frequently.
It is interesting to note that many latter‐day authors have attributed the introduction
of the term ‘fuel cell’ to Ludwig Mond and Charles Langer in their description of a new
form of gas battery in 1889. Remarkably, however, there is no mention of ‘fuel cell’ in this
communication. Other claims that William W. Jacques, in reporting his experiments to
produce electricity from coal, coined the name are equally ill founded. A. J. Allmand in his
book The Principles of Applied Electrochemistry, published in 1912, appears to attribute
the appellation ‘fuel cell’ to the Nobel Laureate Friedrich Wilhelm Ostwald in 1894.
Grove concluded his short paper in 1842 with the following modest entreaty:
‘Many other notions crowd upon my mind, but I have occupied sufficient space
and must leave them for the present, hoping that other experimenters will think
the subject worth pursuing.’
Unfortunately, however, the invention of the first internal combustion engine to
become commercially successful by Jean Joseph Étienne Lenoir in 1859, coupled
ironically with Faraday’s earlier discovery of electromagnetic force, diverted the course
of electricity generation from electrochemical to electromagnetic methods. As a result,
the fuel cell became merely an object of scientific curiosity during much of the next
half‐century. Meanwhile, knowledge of electrochemical conversion and storage of
energy progressed largely through the development of battery technologies.
In 1894, a well‐documented criticism against heat engines was expressed by Friedrich
Ostwald, who drew attention to the poor efficiency and polluting problems associated
with producing electrical power via the combustion of fossil fuels rather than by direct
electrochemical oxidation. A fuel cell is inherently a more thermodynamically efficient
device since, unlike an engine in which heat is converted to mechanical work, the cell is
not subject to the rules of the Carnot cycle. By virtue of this cycle, the efficiency of
a thermal engine is always lowered to a value far below 100%, as determined by the
difference between the temperature at which heat is taken in by the working fluid and
the temperature at which it is rejected. On this basis, Ostwald advocated that:
‘The path which will help to solve this biggest technical problem of all, this path
must be found by the electrochemistry. If we have a galvanic element which directly
delivers electrical power from coal and oxygen, […] we are facing a technical
revolution that must push back the one of the invention of the steam engine.
Imagine how […] the appearance of our industrial places will change! No more smoke,
no more soot, no more steam engine, even no more fire, […] since fire will now only
be needed for the few processes that cannot be accomplished electrically, and those
will daily diminish. […] Until this task shall be tackled, some time will pass by.’
5
6
Fuel Cell Systems Explained
Regrettably, Ostwald was proven to be correct as regards his closing prediction for
although attempts were made at the turn of the century to develop fuel cells that
could convert coal or carbon into electricity (for instance, the work of William W.
Jacques in the United States), the need for an expensive platinum catalyst and its
poisoning by carbon monoxide formed during the coal gasification limited cell
affordability, usefulness and lifetime. Consequently, interest in such ‘direct carbon
fuel cells’ dwindled.
In the 1930s, Emil Bauer and H. Preis in Switzerland experimented with solid oxide
fuel cells (SOFCs). Given the limitations of solid oxides at that time (i.e., poor electrical
conductivity and chemical stability), G.H.J. Broers and J.A.A. Ketelaar in the late 1950s
turned to the use of fused salts as electrolytes. The work gave birth to the molten
carbonate fuel cell (MCFC), which eventually became one of the main types of fuel cell
in commercial production.
The renaissance of the fuel‐cell concept in the 20th century can be attributed
largely to the work of Englishman F.T. (Tom) Bacon. He was an engineer by profession
and thus appreciated the many potential advantages of the fuel cell over both the
internal combustion engine and the steam turbine as a source of electrical power.
His interest in fuel cells dated as far back as 1932, and he ploughed a lone furrow,
with little support or backing, but showed enormous dedication to the challenge of
developing practical cells. Early in his career, Bacon elected to study the alkaline‐
electrolyte fuel cell (AFC), which used nickel‐based electrodes, in the belief that
platinum‐group electrocatalysts would never become commercially viable. In
addition, it was known that the oxygen electrode is more readily reversible in
alkaline solution than in acid. This choice of electrolyte and electrodes necessitated
operating the cell at moderate temperatures (100–200°C) and high gas pressures.
Bacon restricted himself to the use of pure hydrogen and oxygen as reactants.
Eventually, in August 1959, he demonstrated the first workable fuel cell — a 40‐cell
system that could produce about 6 kW of power, which was sufficient to run a forklift
truck and to operate a welding machine as well as a circular saw.
A major opportunity to apply fuel cells arose in the early 1960s with the advent of
space exploration. In the United States, fuel cells were first employed to provide
spacecraft power during the fifth mission of Project Gemini. Batteries had been
employed for this purpose in the four earlier flights, as well as in those conducted in
the preceding Project Mercury. This switch in technology was undertaken because
payload mass is a critical parameter for rocket‐launched satellites, and it was judged
that fuel cells, complete with gas supplies, would weigh less than batteries. Moreover,
the objective of Project Gemini was to evolve techniques for advanced space travel —
notably, the extravehicular activity and the orbital manoeuvres (rendezvous, docking, etc.) required for the moon landing planned in the following Project Apollo.
Thus, lunar flights demand a source of power of longer duration than that available
from batteries.
A proton‐exchange membrane fuel cell (PEMFC) system manufactured by the
General Electric Company was adopted for the Gemini missions (two modules, each
with a maximum power of about 1 kW), but this was replaced in Project Apollo by an
AFC of circulating electrolyte design, as pioneered by Bacon and developed by the Pratt
and Whitney Aircraft Company (later the United Technologies Corporation). Both
Introducing Fuel Cells
types of system were fuelled by hydrogen and oxygen from cryogenic tanks. The AFC
could supply 1.5 kW of continuous power, and its in‐flight performance during all 18
Apollo missions was exemplary. In the 1970s, International Fuel Cells (a division of
United Technologies Corporation) produced an improved AFC for the Space Shuttle
orbiter that delivered eight times more power than the Apollo version and weighed
18 kg less. The system provided all of the electricity, as well as drinking water, when the
Space Shuttle was in flight.
The successful exploitation of fuel cells in the space programme drove research
activity worldwide during the 1970s to develop systems that would generate power
with high efficiency and low emissions for terrestrial applications. Research was
stimulated further by the hiatus in the global oil supply in 1974. What followed was
the emergence of various national initiatives on fuel‐cell development. In the United
States, demonstrations of phosphoric acid fuel cell (PAFC) technology by the
American Gas Association led to a Notice of Market Opportunities (NOMO) initiative.
This activity, in turn, renewed interest in the MCFC by US researchers, and in the
mid‐1980s, national research and development programmes were established in
Japan and Europe. Renewed interest in the PEMFC was championed in the late 1980s
by Geoffrey Ballard, a Canadian pioneer, who saw the potential for the technology to
replace internal combustion engines. Since then, this system has been the subject
of much advancement for a variety of applications, so much so that it merits two
chapters in this book.
1.2
Fuel‐Cell Basics
To understand how the reaction between hydrogen and oxygen produces an electric
current, and where the electrons are released, it is necessary to consider the reaction
that takes place at each electrode. The reactions vary for different types of fuel cell, but
it is convenient to start with a cell based around an acid electrolyte, not only because
this system was used by Grove but also because it is the simplest and still the most
chosen for commercial applications.
At the anode of an acid fuel cell, hydrogen is oxidized and thereby releases electrons
and creates H+ ions, as expressed by:
2H 2
4H
4e
(1.1)
This reaction also releases energy in the form of heat.
At the cathode, oxygen reacts with electrons taken from the electrode, and H+ ions
from the electrolyte, to form water, i.e.,
O2 4 e
4H
2H 2 O
(1.2)
Thus the overall cell reaction is:
2H 2 O 2
2H2O heat
(1.3)
7
8
Fuel Cell Systems Explained
Clearly, for both the electrode reactions to proceed continuously, electrons produced
at the negative electrode must pass through an electrical circuit to the positive. Also,
H+ ions must pass through the electrolyte solution — an acid is a fluid with free H+ ions
and so serves this purpose very well. Certain polymers and ceramic materials can also
be made to contain mobile H+ ions. These materials are commonly called ‘proton‐
exchange membranes’, as an H+ ion is also known as a proton. The PEMFC is examined
in detail in Chapter 4.
The cell reaction (1.3) shows that two hydrogen molecules will be needed for each
oxygen molecule if the system is to be kept in balance. The operating principle is
illustrated in Figure 1.3.
In a fuel cell with an alkaline electrolyte (AFC), the overall reaction of hydrogen
oxidation is the same, but the reactions at each electrode are different. In an alkaline
solution, hydroxyl (OH−) ions are available and mobile. At the anode, these ions react
with hydrogen to release electrons and energy (heat) together with the production
of water:
2H2 4OH
4 H2O 4 e
(1.4)
At the cathode, oxygen reacts with electrons taken from the electrode, and water in
the electrolyte and thereby forms new OH− ions:
O2 4 e
2H 2 O
4OH
(1.5)
Comparing equations (1.4) and (1.5) shows that, as with an acid electrolyte, twice as
much hydrogen is required compared with oxygen. The operating principle of the AFC
is presented in Figure 1.4.
There are many other types of fuel cell, each distinguished by its electrolyte and the
reactions that take place on the electrodes. The different systems are described in detail
in the following chapters.
Hydrogen fuel
– Anode
4H+
2H2
+
4e–
Load
e.g., electric
motor
H+ Ions through electrolyte
+
Cathode O2
+ 4e–
+
4H+
Oxygen, usually from the air
2H2O
Electrons flow round
the external circuit
Figure 1.3 Electrode reactions and charge flow for fuel cell with an acid electrolyte. Note that
although the negative electrons flow from the anode to cathode, the ‘conventional positive current’
flows from cathode to anode.
Introducing Fuel Cells
Hydrogen fuel
–
Anode 2H2
+
4OH–
4H2O
+
4e–
Load
e.g., electric
motor
OH– Ions through electrolyte
+
Cathode O2
+
4e–
+
4OH–
2H2O
Electrons flow round
the external circuit
Oxygen, usually from the air
Figure 1.4 Electrode reactions and charge flow for a fuel cell with an alkaline electrolyte. Electrons
flow from negative anode to positive cathode, but ‘conventional positive current’ flows from cathode
to anode.
1.3
Electrode Reaction Rates
The oxidation of hydrogen at the negative electrode liberates chemical energy. It does
not follow, however, that the reaction proceeds at an unlimited rate; rather, it has
the ‘classical’ energy form of most chemical reactions, as shown in Figure 1.5. The
schematic represents the fact that some energy must be used to excite the atoms or
molecules sufficiently to start the chemical reaction — the so‐called ‘activation energy’.
This energy can be in the form of heat, electromagnetic radiation or electrical energy.
In visual terms, the activation energy helps the reactant to overcome an ‘energy hill’,
Energy
Activation
energy
Energy
released
Stage of reaction
Figure 1.5 Classical energy diagram for a simple exothermic chemical reaction.
9
10
Fuel Cell Systems Explained
and once the reaction starts, everything rolls downhill. Thus, if the probability of an
atom or molecule having sufficient energy is low, then the reaction will only proceed
slowly. This is indeed the case for fuel‐cell reactions, unless very high temperatures
are employed.
The three main ways of dealing with the slow reaction rates are to (i) use catalysts,
(ii) raise the temperature and (iii) increase the electrode area. Whereas the first two
options can be applied to any chemical reaction, the electrode area has a special
significance for electrochemical cells. The electrochemical reactions take place at the
location where the gas molecules (hydrogen or oxygen) meet the solid electrode and
the electrolyte (whether solid or liquid). The point at which this occurs is often
referred to as the ‘three‐phase boundary/junction’ or the ‘triple‐phase boundary/
junction’ (v.s.).
Clearly, the rate at which either electrode reaction proceeds will be proportional to
the area of the respective electrode. Indeed, electrode area is such an important issue
that the performance of fuel cells is usually quoted in terms of the current per cm2.
Nevertheless, the geometric area (length × width) is not the only issue. The electrode
is made highly porous so as to provide a great increase in the ‘effective’ surface area
for the electrochemical reactions. The surface area of electrodes in modern fuel cells,
such as that shown in Figure 1.6, can be two to three orders of magnitude greater
than the geometric area. The electrodes may also have to incorporate a catalyst
and endure high temperatures in a corrosive environment; catalysts are discussed
in Chapter 3.
Figure 1.6 Transmission electron
microscope image of a fuel‐cell
catalyst. The black spots are the
catalyst particles that are finely
divided over a carbon support. The
structure clearly has a large surface
area. (Source: Courtesy of Johnson
Matthey Plc.)
75 nm
Introducing Fuel Cells
1.4
Stack Design
Because a fuel cell functions at a low voltage (i.e., well below 1 V), it is customary to
build up the voltage to the desired level by electrically connecting cells in series to form
a ‘stack’. There are a number of different designs of fuel cell, but in each case the unit cell
has certain components in common. These are as follows:
●
●
●
●
●
An electrolyte medium that conducts ions. This may be a porous solid that contains
a liquid electrolyte (acid, alkali or fused salt) or a thin solid membrane that may be a
polymer or a ceramic. The membrane must be an electronic insulator as well as a
good ionic conductor and must be stable under both strong oxidizing and strong
reducing conditions.
A negative fuel electrode (anode) that incorporates an electrocatalyst, which is
dispersed on an electronically conducting material. The electrode is fabricated so
that the electrocatalyst, the electrolyte and the fuel come into simultaneous contact
at a three‐phase boundary (v.s.).
A positive electrode (cathode), also with a triple‐point electrocatalyst, at which the
incoming oxygen (either alone or in air) is reduced by uptake of electrons from the
external circuit.
A means of electrically connecting individual cells together. The design of interconnector depends on the geometry adopted for the cells.
Seals that keep the gases apart and also prevent cell‐to‐cell seepage of liquid electrolyte,
which otherwise would give rise to partial short-circuits.
A stack also has current-collectors that are located at the two ends of the stack and are
connected by end‐plate assemblies.
Historically, the flat plate is by far the preferred geometry for fuel cells, and one way
of assembling such cells in series is to connect the edge of each negative electrode to the
positive of the next cell through the string, as illustrated in Figure 1.7. (For simplicity,
the diagram ignores the difficulty of supplying gas to the electrodes.) The problem with
this method, however, is that the electrons have to flow across the face of the electrode
to the current collection point at the edge. The electrodes might be quite good
conductors, but if each cell is only operating at about 0.7 V, even a small voltage drop
can be significant. Consequently, this type of stack design is not used unless the current
flows are very low, the electrodes are particularly good conductors and/or the dimensions
of the stack are small.
A much better method of cell interconnection for planar fuel cells is to use a ‘bipolar
plate’. This is an electrically conducting plate that contacts the surfaces of the positive
electrode of one cell and the negative electrode of the next cell (hence the term ‘bipolar’).
At the same time, the bipolar plate serves as a means of feeding oxygen to the negative
anode and fuel gas to the positive cathode of the adjacent cells. This is achieved by
having channels machined or moulded on either side of the plate along which the gases
can flow and the products, i.e., pure water in the case of hydrogen fuel, can exit. Various
designs of channel geometry have been proposed to maximize the access of gases
and the removal of water, e.g., pin‐type, series–parallel, serpentine, integrated and
interdigitated flow-fields. The different types are described in later chapters when
considering the stacking arrangement of each type of fuel cell. The arrangement of the
11
12
Fuel Cell Systems Explained
Load
Oxygen fed to
each cathode
Hydrogen
fed to each
anode
Cathode
Electrolyte
Anode
For reactions in this part
the electrons have to pass all
along the face of the electrode
Figure 1.7 Simple edge connection of three‐planar fuel cells in series. When the electrolyte is a
membrane, the cathode–electrolyte–anode unit is generally known as a membrane–electrode
assembly (MEA).
channels (also known as the ‘flow-field’) leads the bipolar plate to be also known as the
flow‐field plate. Bipolar plates must also be relatively impermeable to gases, sufficiently
strong to withstand stack assembly and easily mass produced. They are made of a good
electronic conductor such as graphite or stainless steel. For transport applications, low
weight and low volume are essential. The method of connecting two plates to a single
cell is illustrated in Figure 1.8; the respective gases are supplied orthogonally.
To connect several cells in series, anode–electrolyte–cathode assemblies have to be
prepared. These are then ‘stacked’ together with bipolar plates placed between each
pair of cells. In the particular arrangement shown in Figure 1.9, the stack has vertical
channels for feeding hydrogen over the anodes and horizontal channels for feeding
oxygen (or air) over the cathodes. The result is a solid block, in which the electric
current passes efficiently more or less straight through the cells, rather than over the
surface of each electrode one after the other.
The electrodes and electrolytes are also well supported, and the whole structure is
clamped together to give a strong and robust device. Although simple in principle, the
design of the bipolar plate has a significant effect on fuel‐cell performance. If the electrical
connection between cells is to be optimized, then the area of contact points should be
as large as possible, but this would mitigate good gas flow over the electrodes. If the
contact points have to be small, at least they should be frequent. This may render
the plate more complex, difficult and expensive to manufacture, as well as fragile.
Ideally, bipolar plates should be as thin as possible so as to minimize both the electrical
resistance between individual cells and the stack size. On the other hand, such an
Introducing Fuel Cells
Anode
Electrolyte
Cathode
Hydrogen fed along
these channels
Negative
connection
Air or oxygen
fed to cathode
Positive
connection
Figure 1.8 Single cell with end-plates for collecting current from the whole face of the adjacent
electrode and applying gases to each electrode.
Hydrogen fed along
these vertical channels
over the anodes
Negative
connection
Positive
connection
Air or oxygen fed
over the cathodes
through these channels
Figure 1.9 A three‐cell stack showing how bipolar plates connect the anode of one cell to the
cathode of its neighbour.
approach would narrow the gas channels and thereby place greater demands on the
pumps for supplying gases. High rates of flow are sometimes required, especially when
using air instead of pure oxygen at the positive electrode. For low‐temperature fuel cells,
the circulating air has to evaporate and carry away the product water. Moreover, in many
13
14
Fuel Cell Systems Explained
cases, additional channels have to pass through the bipolar plate to carry a cooling fluid.
Some further challenges for the bipolar plate are considered in the next section.
1.5
Gas Supply and Cooling
The arrangement given in Figure 1.9 has been simplified to show the basic principle of
the bipolar plate. In practice, however, the twin problems of gas supply and preventing
leaks mean that the design is somewhat more complex.
Because the electrodes must be porous (to permit the access of gas), they allow leakage
of the gas through their edges. Consequently, the edges must be sealed. Sometimes this
is done by making the electrolyte compartment slightly larger than one, or both, of the
electrodes and fitting a gasket around each electrode, as presented in Figure 1.10. Such
assemblies can then be made into a stack in which the fuel and oxygen can then be
supplied to the electrodes using the external manifolds as shown disassembled in
Figure 1.11. With this arrangement, the hydrogen should only come into contact with
the anodes as it is fed vertically through the fuel‐cell stack. Similarly, the oxygen (or air)
fed horizontally through the stack should only contact the cathodes and certainly not
the edges of the anodes. Such would not be the case for the basic design illustrated in
Figure 1.9.
The externally manifolded design suffers from two major disadvantages. The first is that
it is difficult to cool the stack. Fuel cells are far from 100% efficiency, and considerable
quantities of heat are generated, as well as electrical power. In practice, the cells in this
type of stack have to be cooled by the reactant air passing over the positive electrodes.
This means that air has to be supplied at a higher rate than that demanded by the
cell chemistry — sometimes the flow is sufficient to cool the cell, but it is wasteful of
energy. The second disadvantage of external manifolding is that there is uneven
pressure over the gasket round the edge of the electrodes, i.e., at the points where there
Edge sealing gasket
Electrolyte
Edge sealing gasket
Anode
Assembly
Cathode
Figure 1.10 The construction of cathode–electrolyte–anode units with edge seals that prevent the
gases leaking in or out through the edges of the porous electrodes.
Introducing Fuel Cells
Manifolds
Cathode – electrolyte – anode
assemblies
Figure 1.11 Three‐cell stack, with external manifolds. Unlike the stack shown Figure 1.9, the
electrodes now have edge seals.
is a channel and the gasket is not pressed firmly onto the electrode. This increases the
probability of leakage of the reactant gases.
‘Internal manifolding’ is a more common stack arrangement and requires a more
complex design of bipolar plate, such as that displayed schematically in Figure 1.12. In
this arrangement, the plates are made larger relative to the electrodes and have extra
channels running through the stack for the delivery of fuel and oxygen to the electrodes.
Holes are carefully positioned to feed the reactants into the channels that run over the
surface of the electrodes. Reactant gases are fed in at the ends of the stack where the
respective positive and negative electrical connections are also made. An example of a
commercial fuel‐cell stack is shown in Figure 1.13.
A stack with internal manifolding can be cooled in various ways. The most practical
method is to circulate a liquid coolant through electrically conductive metal plates that
are inserted between groups of cells. In this passive approach, the heat within the plane
of the plate must be conducted out to one or more of the edges of the plate for transfer
to a heat-exchanger external to the fuel‐cell stack. Alternatively the bipolar plates
themselves can be made thicker and machined to incorporate extra channels that allow
passage of cooling air or water. The preferred cooling method varies greatly with the
type of fuel cell and is addressed in later chapters.
From the foregoing discussion, it should be apparent that the bipolar plate is a key
component of a fuel‐cell stack. As well as being a fairly intricate item to manufacture,
the choice of material for its construction raises issues. For low‐temperature fuel
15
16
Fuel Cell Systems Explained
Air supplied
through here
Air removed
through here
Hydrogen removed
through here
Channel for
distributing air
over cathode
Hydrogen supplied
through here
Channel for supplying
hydrogen to surface of anode
Figure 1.12 Internal manifolding. A more complex bipolar plate allows reactant gases to be fed to
electrodes through internal tubes. (Source: Courtesy of Ballard Power Systems.)
Figure 1.13 A 96‐cell, water‐cooled PEMFC stack that produces up to 8.4 kW and weighs 1.4 kg.
(Source: Courtesy of Proton Motor GmbH.)
cells, graphite was one of the first materials to be employed, but it is difficult to work
and brittle and, consequently, has now largely been replaced by various carbon
composite materials. Stainless steel can also be used, but it will corrode in some types
of fuel cell. Ceramic materials have found application in fuel cells that operate at high
temperatures. The bipolar plate nearly always is a major contributor to the capital
cost of a fuel cell.
Introducing Fuel Cells
1.6
Principal Technologies
Setting aside practical issues such as manufacturing and materials costs, the two
fundamental technical problems with fuel cells are:
●
●
The slow reaction rates, particularly for the oxygen reduction reaction, which lead to
low levels of current and power.
The fact that hydrogen is not a readily available fuel1.
To address these problems, many different types of fuel cell have been developed and
tested. The systems are usually distinguished by the electrolyte that is used and the
operating temperature, though there are always other important differences as well.
There are six principal types of fuel cell, namely:
●
●
●
Low temperature (50–150°C): alkaline electrolyte (AFC), proton‐exchange membrane (PEMFC), direct methanol (DMFC) and other liquid‐fed fuel cells.
Medium temperature (around 200°C): PAFC.
High temperature (600–1000°C): molten carbonate (MCFC) and SOFC.
Some operational data on each type are given in Table 1.1. There are other less well‐
known types such as the direct borohydride (DBFC) and direct carbon fuel cells (DCFC);
the former operates at low temperatures and the latter at high temperatures.
Table 1.1 Principal types of fuel cell.
Fuel cell type
Operating
Mobile ion temperature (°C) Fuel
Alkaline (AFC)
OH–
50–200
Pure H2
Space vehicles, e.g.,
Apollo, Shuttle
Proton‐exchange
membrane
(PEMFC)
H+
30–100 + a
Pure H2
Vehicles and mobile
applications, and for
lower power CHP systems
Direct methanol
(DMFC)
H+
20–90
Methanol
Portable electronic
systems of low power,
running for long times
Phosphoric acid
(PAFC)
H+
~220
H2, (low S, low CO,
tolerant to CO2)
Large numbers of 200‐kW
CHP systems in use
~650
H2, various
hydrocarbon fuels
(no S)
Medium‐ to large‐scale
CHP systems, up to MW
capacity
500–1000
Impure H2, variety
All sizes of CHP systems,
of hydrocarbon fuels 2 kW to multi MW
Molten carbonate CO32−
(MCFC)
Solid oxide
(SOFC)
O2−
Applications and notes
CHP, combined heat and power.
a) New electrolyte materials as described in Chapter 4 are enabling higher operating temperatures for the
PEMFC.
1 Although hydrogen is preferred for most types of fuel cell, other fuels can be used for some technologies.
For example, methanol is employed in the direct methanol fuel cell (DMFC) and carbon as the fuel in the
direct carbon fuel cell (DCFC).
17
18
Fuel Cell Systems Explained
To date, the PEMFC has proved to be the most successful commercially. The electrolyte
is a solid polymer, in which protons are mobile. The chemistry is the same as that shown
Figure 1.3 for an acid‐electrolyte system. The PEMFC runs at relatively low temperatures,
so the problem of slow reaction rates is addressed by using sophisticated catalysts
and electrodes. Platinum has been the preferred catalyst. It is an expensive metal
but, through improvements in materials, only minute amounts are now required.
Consequently, in modern PEMFC designs, the platinum makes a relatively small
contribution to the total cost of the fuel‐cell system. More recent research suggests that
in some cases platinum can be eliminated from the catalyst. Further discussion of the
PEMFC is given in Chapter 4. The PEMFC has to be fuelled with hydrogen of high
purity, and methods for meeting this requirement are discussed in Chapter 10.
The DMFC is a variant of the PEMFC. The technology differs from the PEMFC only
in that methanol in its native liquid form is used as fuel. Other liquid fuels such as
ethanol and formic acid may also be viable for some applications. Unfortunately, most
of these liquid‐fuelled cells produce very low levels of power, but, even with this
limitation, there are many potential applications for such devices in the rapidly growing
area of portable electronics devices. Such cells, for the foreseeable future at least, will
remain low‐power units and will therefore suit applications that require slow and steady
consumption of electricity over long periods.
As mentioned earlier, an AFC system was chosen for the Apollo and Space Shuttle
orbiter craft. The problem of slow reaction rate was overcome by using highly porous
electrodes, with a platinum catalyst, and sometimes by operating at quite high
pressures. Although some historically important AFCs have been run at about 200°C,
the systems usually operate below 100°C. Unfortunately, the AFC is susceptible to
poisoning by the carbon dioxide in the atmosphere. Thus the air and fuel supplies must
be free from this gas, or else pure oxygen and hydrogen must be supplied.
The PAFC was the first type of fuel cell to reach commercialization and the technology
enjoyed a reasonable degree of widespread terrestrial use in the period 1980–2000.
Many 200‐kW systems, manufactured by the International Fuel Cells Corporation, were
installed in the United States and Europe. Other systems were produced by Japanese
companies. In the PAFC, porous electrodes, platinum catalysts and a moderately high
temperature (~220°C) help to boost the reaction rate to a reasonable level. Such PAFC
systems were fuelled with natural gas, which is converted to hydrogen within the fuel‐
cell system by steam reforming. The required equipment for steam reforming unfortunately adds considerably to the costs, complexity and size of the fuel‐cell system.
Nevertheless, PAFC systems have demonstrated good performance in the field, for
instance, units have run for periods in excess of 12 months without any maintenance
that has required shutdown or human intervention. A typical installation of a 400 kW
PAFC system is shown in Figure 1.14.
The most common form of SOFC operates in the region of 600–1000°C. These high
temperatures permit high reaction rates to be achieved without the need for expensive
platinum catalysts. At these elevated temperatures, fuels such as natural gas can be
used directly (internally reformed) within the fuel cell without the need for a separate
processing unit. The SOFC thus addresses the aforementioned key problems (viz. slow
reaction rates and hydrogen supply) and takes full advantage of the inherent simplicity
of the fuel‐cell concept. Nevertheless, SOFCs are made from thin ceramic materials
that are difficult to handle and therefore are expensive to manufacture. In addition,
Introducing Fuel Cells
Figure 1.14 Phosphoric acid fuel cell for stationary power‐plant applications (Source: Creative
commons – Courtesy of UTC.)
a large amount of extra equipment is needed to make a full SOFC system, e.g., air and
fuel preheaters, heat-exchangers and pumps. Also the cooling system is more complex
than for low‐temperature fuel cells. Care also has to be taken during start‐up and
shutdown of SOFC systems, on account of the intrinsic fragile nature of the ceramic
materials in the stacks.
The MCFC has an interesting and distinguishing feature in that it requires carbon
dioxide to be fed to the positive electrode, as well as oxygen. This is usually achieved
by recycling some of the exhaust gas from the anode to the cathode inlet. The high
temperature means that a good reaction rate is achieved with a comparatively inexpensive
catalyst — nickel. Like the SOFC, an MCFC system can be fuelled directly with gases,
such as methane and coal gas (a mixture of hydrogen and carbon monoxide), without
the need for an external reformer. This advantage for the MCFC is somewhat offset,
however, by the nature of the electrolyte, namely, a hot and corrosive molten mixture of
lithium, potassium and sodium carbonates.
1.7 Mechanically Rechargeable Batteries
and Other Fuel Cells
At the start of this book, a fuel cell was defined as an electrochemical device that
converts a fuel to electrical energy (and heat) continuously, as long as reactants are
supplied to its electrodes. The implication is that neither the electrodes nor the electrolyte
is consumed by operation of the cell. Of course, in all fuel cells the electrodes and
19
20
Fuel Cell Systems Explained
electrolytes are degraded and subject to ‘wear and tear’ during service. The first two
technologies under consideration in this section are often misleadingly described as
fuel cells and employ electrodes that are entirely consumed during use.
1.7.1
Metal–Air Cells
The most common type of cell of this category is the zinc–air battery, though aluminium–
air and magnesium–air cells have been produced commercially. In all cases, the basic
operation is the same.
At the negative electrode, the metal reacts with hydroxyl ions in an alkaline electrolyte
to form the metal oxide or hydroxide. For example, the reaction with a zinc fuel is given by:
Zn 2OH
ZnO H2O 2e
(1.6)
The electrons thus released pass round the external electric circuit to the air electrode
where they are available for the reaction between water and oxygen to form more
hydroxyl ions. Thus at the air electrode the reaction is exactly the same as equation (1.5)
for the AFC. Cells using a salt solution (e.g., seawater) as the electrolyte solution also
work reasonably well when using aluminium or magnesium as the fuel.
Metal–air cells have a very high specific energy (Wh kg−1). Zinc–air batteries are
employed widely in devices that require long running times at low currents, such as
hearing aids. Some interest has also been shown in the development of units with higher
power for application in electric vehicles. Such systems can also be ‘refuelled’ by
replenishing the metal consumed at the negative electrode — which is why the technology
is sometimes promoted as a ‘fuel cell’. This claim is also supported by the fact that the
reaction at the positive electrode is exactly the same as for a fuel cell, and indeed the
same electrodes can be used. It should be noted, however, that removal of the metal
oxide will also necessitate renewal of the electrolyte solution. Thus, the metal–air systems
cannot properly be described as fuel cells and are best classified as ‘mechanically
rechargeable batteries’.
1.7.2
Redox Flow Cells
Another type of electrochemical power source that is sometimes taken to be a fuel cell
is the ‘redox flow cell’ (or ‘flow cell’); a multicell unit is usually referred to as a ‘flow
battery’. It is useful at this point to define two types of flow cell, as several different
chemistries are under development:
1) Flow batteries, in which there is a decoupling of cell power and cell capacity, e.g., the
bromine–polysulfide cell and the vanadium redox cell.
2) Hybrid flow batteries, in which there is no decoupling of cell power and cell capacity,
e.g., the zinc–bromine battery.
The first category is different from all other fuel cells in that the oxidant is not air, and
therefore it cannot be said that the fuel is ‘combusted’. In this type of cell, there is one
reactant (which can be called the fuel) that is oxidized and a complementary reactant
that serves as the oxidant. These are removed from the electrode compartments when
the cell is being charged and stored in tanks. The capacity of such cells can thus be very
large. Discharge is undertaken by resupplying the reactants to the electrodes.
Introducing Fuel Cells
Two flow batteries have been the subject of much research, namely, the sodium‐
bromide–sodium‐polysulfide cell and the vanadium redox cell. The former cell was
introduced in the 1990s by Regenesys Technologies Limited in the United Kingdom.
After a utility‐scale demonstration at a power station by the National Power in
Cambridgeshire, United Kingdom, the development was taken over by RWE and
subsequently by Prudent Energy to complement its own work on a vanadium battery.
No further studies or trials of the Regenesys have been reported.
The vanadium redox battery was pioneered in the 1980s at the University of New South
Wales in Sydney, Australia, and the Japanese Electrotechnical Laboratory. The operating
principle of the system is illustrated in Figure 1.15. The two reactants are flowing aqueous
solutions of vanadium sulfate and the electrode reactions are as follows.
At the positive electrode:
VO2
Discharge
2H
e
Charge
VO2
(1.7)
H2O
At the negative electrode:
V2
Discharge
Charge
V3
(1.8)
e
Thus, in the charged state, the positive‐electrolyte loop contains a solution of V5+ and
the negative loop contains a solution of V2+. On discharging, the former solution is
reduced to V4+ and the latter is oxidized to V3+. The difference in the oxidation state of
e–
e–
Load or
power source
e
e
d
ro
El
Tank
V2+/V3+
e
an
d
ro
br
t
ec
t
ec
em
Charge
V2+
V3+
M
Charge
V4+
El
Tank
V5+/ V 4+
V5+
+
–
Discharge
Ions
Pump
Discharge
Pump
Figure 1.15 Operating principle of the vanadium redox battery.
21
22
Fuel Cell Systems Explained
vanadium in the two reactant solutions produces 1.2–1.6 V across the membrane, as
determined by the electrolyte solution, temperature and state-of-charge. Regeneration
takes place by reversing the flow of the solutions and applying a potential across the cell
to restore the original oxidation states in the solutions.
It can easily be seen that (i) this is a reversible cell and (ii) the capacity of the cell (e.g.
as measured in kWh) is determined by the amount of liquid pumped, i.e., the size of the
storage tanks, and not by the dimensions of the electrodes as would be the case in a
normal battery. Furthermore, the more the cells and the faster the flow of electrolyte
solutions, the higher is the power rating. This approach enables economies of scale in
both manufacturing and energy–power capacity.
The vanadium redox cell shares many characteristics with now‐abandoned Regenesys.
Numerous companies and organizations have been involved in funding and developing
the vanadium technology, and several large field trials have been conducted around the
world. Research and development is continuing.
In the hybrid form of flow cell, one or more of the electroactive components are
deposited as a solid layer. Consequently, the system may be viewed as a combination of
one battery electrode and one fuel‐cell electrode. The zinc–bromine system is the best‐
known example of such technology. A modern version developed by Redflow Limited,
an Australian‐based company, is shown in Figure 1.16. As with the vanadium redox cell,
the zinc–bromine cell is comprised of two fluids that pass carbon‐plastic electrodes
that are each placed in a half‐cell either side of a microporous polyolefin membrane.
During discharge, zinc and bromine combine into zinc bromide and thereby generate
1.8 V across each cell. During charge, metallic zinc will be drawn out of solution and
deposited (plated) as a thin film on one side of the negative electrode. Meanwhile,
bromine evolves as a dilute solution at the positive electrode on the other side of the
membrane. Because bromine is a highly volatile and reactive liquid, it is complexed with an
organic reagent to form a poly‐bromo compound, which is an oil and is immiscible with
the aqueous electrolyte solution. The oil sinks down to the bottom of the electrolytic
Figure 1.16 Redox zinc–bromine battery. (Source: Courtesy of Redflow Pty Ltd.)
Introducing Fuel Cells
tank and is separated and stored in a special compartment in the external reservoir
of the positive electrode until required again for discharge. The capacity of the cell is
limited by the amount of zinc that can be plated on the negative electrode.
1.7.3
Biological Fuel Cells
Finally, it should be noted that, although not yet a principal technology, the biological
fuel cell is attracting interest as a long‐term prospect. The cell would normally operate
with an organic fuel, such as methanol or ethanol. The distinctive ‘biological’ aspect is
that the electrode reactions are promoted by enzymes present in microbes, rather than
by conventional ‘chemical’ catalysts such as platinum. Hence, these systems — also
known as ‘microbial fuel cells (MFCs)’ — replicate nature in the way that energy is
derived from organic fuels. Biological or microbial fuel cells should be distinguished
from biological methods for generating hydrogen, which is then used in a conventional
fuel cell. Such methods of hydrogen production are discussed in Chapter 10. Research
into advanced microfluidics, new bacterial strains, more robust separator membranes
and efficient electrodes is the key to unlocking the potential of MFCs.
1.8
Balance‐of‐Plant Components
It should be evident that a practical fuel‐cell system requires not only a readily available
fuel but also a means of cooling the stack, an ability to employ the heat produced to do
useful work and an application for the direct current (dc) power that is produced by
the stack(s). For a fuel‐cell stack to function effectively, various other components are
necessary. The exact composition of this so‐called balance-of-plant depends on the type of
fuel cell, the available fuel and its purity and the desired outputs of electricity and heat.
Typical auxiliary subsystems are:- (i) fuel clean‐up processor, e.g., for sulfur removal —
so‐called desulfurization; (ii) steam reformer and shift reactor for the fuel; (iii) carbon
dioxide separator; (iv) humidifier; (v) fuel and air delivery units; (vi) power‐conditioning
equipment, e.g., for inverting dc to alternating current (ac) and then transforming to
line voltage; (vii) facilities for the management of heat and water; (viii) overall control
and safety systems; and (ix) thermal insulation and packaging. Individual components
include fuel storage tanks and pumps, compressors, pressure regulators and control
valves, fuel and/or air pre‐heaters, heat-exchangers and radiators, voltage regulators,
motors and batteries (to provide power for pumps on start‐up). These important
subsystem issues are described in much more detail in Chapter 12.
The requirements for a fuel‐cell system for a stationary power application and a
vehicle are very different. In a stationary power plant system, such as shown in
Figure 1.14, the fuel‐cell stack is, in terms of size, a small part of the installation that is
dominated by the fuel and heat‐processing systems and the power‐conditioning
equipment. This will nearly always be the case for combined heat and power (CHP)
facilities that run on a conventional fuel such as natural gas.
By contrast, a fuel‐cell power source for a car is shown in Figure 1.17. The unit operates
on gaseous hydrogen fuel that is stored on the vehicle, and the waste heat is only
used to warm the car interior. The fuel‐cell stack occupies the bulk of the compartment
23
24
Fuel Cell Systems Explained
Figure 1.17 Hyundai fuel‐cell system located under the car hood. (Source: Courtesy of Hyundai Motor
Company, Australia.)
that would normally be filled with an internal combustion engine (ICE). Other components
of a hydrogen fuel‐cell ‘engine’ in a vehicle, i.e., pumps, humidifier, power electronics
and compressor, are generally much less bulky than those of a CHP system.
1.9
Fuel‐Cell Systems: Key Parameters
To compare the performance of fuel‐cell systems with each other and with other electric
power generators, some key operating parameters must be considered. For electrodes
and electrolytes, the key criterion is the current per unit area, which is always known as
the ‘current density’ and usually expressed in terms of mA cm−2, except in the United
States where A ft−2 is frequently adopted (the two units are quite similar, i.e.,
1.0 mA cm−2 = 0.8 A ft−2). The current density should be reported at a specific operating
voltage, typically about 0.6 or 0.7 V. The values for current density and selected voltage
can then be multiplied to give the power per unit area, in mW cm−2. A note of caution
should be made here, namely, that electrodes frequently do not ‘scale up’ properly. That
is, if the area is doubled the current will often not double. The reasons for this are varied
but generally relate to issues such as the even delivery of reactants to, and removal of
products from, the entire face of the electrode.
Specific power (kW kg−1) and power density (kW m−3 or kW L−1) are key ‘figures of
merit’ for comparing electrical generators. Note that whereas power is measured in kW,
energy is simply power delivered over a certain period of time and is measured in kWh.
The capital cost of a fuel‐cell system is obviously an important parameter and is usually
quoted in US$ per kW for ease of comparison.
The lifetime of a fuel cell is rather difficult to specify. Standard engineering measures
such as ‘mean time between failures’ (MTBF) are not entirely applicable given that the
performance of a fuel cell always gradually deteriorates and the power drops fairly
Introducing Fuel Cells
steadily with time as the electrodes and the electrolyte solution both age. The degradation
of a fuel cell is sometimes reported as a decline in cell voltage, given in units of mV per
1000 h. Formally, the life of a fuel cell is considered to be over when it can no longer
deliver the rated power, e.g., when a 10‐kW fuel cell can no longer deliver 10 kW. It
should be noted that, when new, a fuel cell may be capable of providing more than the
rated power, e.g., an extra 25% is not unusual.
The remaining fuel‐cell characteristic of key importance is the efficiency, i.e., the
electrical energy delivered by the system compared with the energy supplied as fuel.
When making comparisons between systems in terms of efficiency, care should be
taken that the data are expressed on the same basis. Efficiency is addressed in Chapter 2.
In the automotive industry, primary issues are the cost per kW and the power density.
In round figures, current ICE technology costs US$10 per kW and delivers 1 kW L−1.
Such a power source should last at least 4000 h, i.e., about 1 h of duty each day for over
10 years. For CHP plant, the capital cost is still important, but a much higher target of
US$1000 per kW is generally accepted. The higher cost is due to the extra balance of
plant that is required and to the fact that the system must have a substantially longer
lifetime. A period of 40 000 h would be a minimum. For stationary power‐generation
systems, the levelized cost of electricity (LCOE) is often used as a measure of performance.
The LCOE is the price at which electricity must be generated from a specific source to
break even over the lifetime of the project. It is an economic assessment of the cost of
the generating system and includes all the costs over its lifetime, namely, capital cost,
operations and maintenance, and cost of fuel. The LCOE enables analysts to compare
the costs of fuel‐cell systems with other forms of power generation.
1.10
Advantages and Applications
For all types of fuel cell, a significant disadvantage or barrier to commercialization is the
capital cost. There are, however, various advantages that feature more or less strongly
for the different systems and lead to fuel cells being attractive for different applications.
These include the following:
●
●
●
●
Efficiency. As explained in Chapter 2, fuel cells are generally more efficient than
piston‐ or turbine‐based combustion engines. A further benefit is that small fuel‐cell
systems can be just as efficient as large ones. This capability opens up a market
opportunity for small‐scale cogeneration (CHP) that cannot be satisfied with turbine‐
or engine‐based systems.
Simplicity. The essentials of a fuel cell involve few, if any, moving parts. This can lead
to highly reliable and long‐lasting systems.
Low emissions. When hydrogen is the fuel, pure water is the by‐product of the main
reaction of the fuel cell. Consequently, the power source is essentially ‘zero emission’.
This is a particularly attractive for vehicle applications, as there is a requirement to
reduce emissions and even eliminate them within cities. Nevertheless, it should be
noted that, at present, emissions of carbon dioxide are nearly always involved in the
production of the hydrogen.
Silence. Fuel cells are very quiet — even those with extensive extra fuel‐processing
equipment. Quietness is very important in both portable‐power applications and for
local power generation via CHP schemes.
25
26
Fuel Cell Systems Explained
Ironically, the fact that hydrogen is the preferred fuel is, in the main, one of the
principal disadvantages of fuel cells. On the other hand, many envisage that as fossil
fuels run out, hydrogen will become a major fuel and energy vector throughout the
world. It could be generated, for example, by electrolysing water using electricity
provided by massive arrays of photovoltaic (solar) cells. Indeed, the so‐called hydrogen
economy may emerge in future decades. In the meantime, it is more likely that ‘hydrogen
energy’ will have only a very small impact globally as it is most economically produced
by the steam reforming of natural gas (see Chapter 10).
In summary, the advantages of fuel cells impact particularly strongly on CHP systems
(both large and small scales) and on mobile power systems ― especially for vehicles
and electronic equipment such as portable computers, mobile telephones and military
communications equipment. A notable feature of the technology is the very wide range
in system sizes, i.e., from a few watts up to several megawatts. In this respect, fuel cells
are unique as energy converters.
Further Reading
Bossel, U, 2000, The Birth of the Fuel Cell 1835‐1845, European Fuel Cell Forum,
Oberrohndorf.
Hoogers, G, 2003, Fuel Cell Technology Handbook, CRC Press, Boca Raton, FL. ISBN
0‐8493‐0877‐1.
27
2
Efficiency and Open‐Circuit Voltage
This chapter examines the efficiency of fuel cells—how it is defined and calculated and
what are the limits. The energy considerations provide information about the open‐
circuit voltage (OCV) of a fuel cell, and the associated formulae yield important
details of the effect on the voltage of factors such as pressure, gas concentration and
temperature.
2.1
Open‐Circuit Voltage: Hydrogen Fuel Cell
The inputs and outputs of energy in a fuel‐cell system are shown schematically in
Figure 2.1. The electrical power and energy output are easily calculated from the following
well‐known formulae:
Power V I
(2.1)
Energy V I t
(2.2)
where V is voltage, I is current and t is time.
By contrast, the energies of the chemical inputs and outputs are less easily defined. In
simple terms, it could be said that the ‘chemical energies’ of the hydrogen, oxygen and
water are involved. The problem is that ‘chemical energy’ can be defined in different
ways—terms such as enthalpy, Helmholtz function and Gibbs free energy are used. In
recent years, the term ‘exergy’ has also become popular,1 and this is particularly useful
when considering the operation of high‐temperature fuel cells. The reader will also
come across older terms such as ‘heating value’ or ‘calorific value’ in the literature.
In the case of fuel cells, it is the ‘Gibbs free energy’ that is fundamentally important.
This can be defined as the energy liberated or absorbed in a reversible process at
constant pressure and constant temperature. Put another way, it is the minimum
thermodynamic work (at constant pressure) required to drive a chemical reaction (or, if
negative, the maximum work that can be done by the reaction). Thus, the Gibbs free
energy is a quantity that can be used to determine if a reaction is thermodynamically
viable or not. The change in free energy, ΔG, in a chemical reaction (i.e., the difference
between the Gibbs free energies of the reactants and products) is given by ΔG = ΔH − TΔS,
1 In thermodynamics, the exergy of a system is the maximum useful work available during a process that
brings the system into equilibrium with its surroundings.
Fuel Cell Systems Explained, Third Edition. Andrew L. Dicks and David A. J. Rand.
© 2018 John Wiley & Sons Ltd. Published 2018 by John Wiley & Sons Ltd.
28
Fuel Cell Systems Explained
Hydrogen
Energy = ?
Electricity
Energy = V × I × t
Fuel
cell
Oxygen
Energy = ?
Figure 2.1 Inputs and outputs of a
fuel cell.
Heat
Water
where ΔH is the change in enthalpy, ΔS is the change in entropy between reactants and
products, and T is the absolute temperature. This expression is known as the ‘Gibbs
equation’.
At this point, it should be remarked that absolute values of the properties of
thermodynamic functions such as enthalpy and entropy are unknown. Only changes
in values caused by parameters such as temperature and pressure can be determined.
It is therefore important to define a baseline for substances, to which the effect of
such variations may be referred. The ‘standard state’ is such a baseline and defines the
standard conditions for temperature and pressure. The International Union of Pure
and Applied Chemistry (IUPAC) has two standards: (i) standard temperature and
pressure, abbreviated as ‘STP’, specifies a temperature of 273.15 K and an absolute
pressure of 100 kPa (1 bar) and (ii) standard ambient temperature and pressure,
abbreviated as ‘SATP’, specifies a temperature of 298.15 K and an absolute pressure of
100 kPa (1 bar).2 It is customary to use the superscript ° to denote that a given quantity
is in its reference state and the subscript f to indicate that a compound is formed from
its elements. Thus G °f is the Gibbs free energy of formation of a compound under
standard state conditions and, therefore, is more frequently referred to as the ‘standard
free energy of formation’. Pure elements are taken to have a free energy of formation
of zero at the reference state. Thus, for an ordinary hydrogen fuel cell operating at
STP, the Gibbs free energy of each reactant (hydrogen and oxygen) is zero—a useful
simplification.
When using thermodynamic functions such as free energy, care should be taken that
the reference states are clearly defined. The standard Gibbs free energy of formation of
a compound, G °f , is the change in Gibbs free energy that accompanies the formation of
1 mol of a substance in its standard state from its constituent elements in their standard
states. Often, the standard state for gases is taken as 298.15 K or 25°C rather than the
recommended 0°C. In most cases, confusion is avoided so long as all of the quantities
are referred to the same standard conditions.
In a fuel cell, it is the change in the Gibbs free energy of formation ΔGf that generates
the electrical energy released by the cell. This change is the difference between the
Gibbs free energy of formation of the products and that of the inputs or reactants,
namely,
Gf
G f products
G f reactants
(2.3)
2 An earlier IUPAC definition of STP in terms of 273.15 K and 1 atm (101.325 kPa) was discontinued in
1982. Reference may also be made to a normal temperature and pressure (NTP), which is usually taken to be
20°C (293.15 K) and 1 atm (101.325 kPa).
Efficiency and Open‐Circuit Voltage
Box 2.1 Molar Mass and the Mole
The ‘mole’ (abbreviation, ‘mol’) is the unit of measurement in the International System of
Units (French: Système international d’unités, SI), which expresses the amount of a given
material. One mole is defined as the number of atoms in precisely 0.012 kg (i.e., 12 g) of
carbon‐12, the most common naturally occurring isotope of the element carbon. This
dimensionless number is equal to approximately 6.022140857 × 1023 and is also called
‘Avogadro’s number’ or the ‘Avogadro constant’. It is represented by the letter NA or L.
The SI unit for molar mass is kg mol−1. For historical reasons, however, molar masses
are almost always expressed in g mol−1. The ‘unified atomic mass unit’ (symbol, u) is
numerically equivalent to 1 g mol−1, i.e., one‐twelfth the mass of one atom of carbon‐12.
(Note that the ‘atomic mass unit’—symbol, amu—without the ‘unified’ prefix is a
technically obsolete unit based on oxygen‐16, but most uses of this term actually refer to
the unified atomic mass unit.) For example, it follows that the molar mass of H2 is 2.0 u,
and therefore 1 g mol of H2 is 2.0 g and 1 kg mol is 2.0 kg. Similarly, the molecular mass of
H2O is 18 u, so 18 g is 1 g mol and 18 kg is 1 kg mol.
A mole of any substance always has the same number of entities (atoms, molecules,
ions, electrons, photons) so that a mole of electrons is 6.022140857 × 1023 electrons. The
charge is NA e−, where e− is 1.60217662 × 10−19 coulombs—the charge on one electron.
This quantity is called the ‘Faraday constant’, which is designated by the letter F and has
the following value:
F N Ae
96 485 coulombs
To make comparisons easier, it is nearly always most convenient to consider these
quantities in their ‘per mole’ form, as discussed in Box 2.1. These can be indicated by a
bar over a lower‐case letter, e.g., ( g f )H2O represents the molar Gibbs free energy of
formation for water.
Consider the basic reaction for the hydrogen–oxygen fuel cell:
2H 2 O 2
(2.4)
2H 2 O
This is equivalent to:
1
O
2 2
H2
(2.5)
H2O
The ‘product’ is 1 mol of H2O, and the ‘reactants’ are 1 mol of H2 and 1/2 mol of O2.
Thus
(2.6)
1
g
g
g
g
f
f HO
2
f H
2
2
f O
2
This equation seems straightforward and simple enough. The Gibbs free energy of
formation is not constant, however, but changes with temperature and state (liquid or
gas). Values of ∆ g f for the basic reaction of the hydrogen fuel cell under a number of
different conditions are listed in Table 2.1. The method used to calculate these data is
outlined in Appendix 1. Note that the values are negative, and therefore, by convention,
this indicates that energy is released by the reaction.
For the hydrogen fuel cell, two electrons pass round the external circuit for each water
molecule produced and each hydrogen molecule used. Thus, for each mole of hydrogen
29
30
Fuel Cell Systems Explained
Table 2.1 ∆ g f for the reaction H2
1
O
2 2
H2O at various temperatures.
(kJ mol−1)
Form of water product
Temperature (°C)
∆g f
Liquid
25
−237.2
Liquid
80
−228.2
Gas
80
−226.1
Gas
100
−225.2
Gas
200
−220.4
Gas
400
−210.3
Gas
600
−199.6
Gas
800
−188.6
Gas
1000
−177.4
consumed, 2NA electrons pass round the external circuit. Given that each electron carries
a unit negative charge (e−), the corresponding charge, in coulombs (C), that flows is
2 N Ae
(2.7)
2F
where F is the Faraday constant or the charge on 1 mol of electrons (see Box 2.1). If V is
the voltage of the fuel cell, then the electrical work, in joules (J), expended in moving
this charge round the circuit is
Electrical work done charge voltage
2FV
(2.8)
If the system is thermodynamically reversible (i.e., it has no energy losses), then the
electrical work done will be equal to the Gibbs free energy released by the fuel‐cell
reaction ∆ g f . Thus:
gf
2 FVr
or Vr
gf
2F
(2.9)
This fundamental equation gives the ‘reversible voltage’, Vr, or ‘OCV’ across the terminals of the cell when there is no net current flow. Under standard conditions, this is
the ‘standard cell voltage’ , Vr°. When the fuel is hydrogen, the reversible voltage under
standard conditions (STP) is 1.229 V at 25°C.
If the cell operates at 200°C, then ∆ g f = −220.4 kJ (from Table 2.1), and thus:
Vr
22.04 103
2 9 6485
1.14 V
(2.10)
Note that this value assumes no ‘irreversibilities’ and that pure hydrogen and oxygen are
supplied at standard pressure (100 kPa). In practice, the voltage would be lower than
this because of the voltage losses discussed in Chapter 3. Some of these irreversibilities
Efficiency and Open‐Circuit Voltage
Box 2.2 Reversible Processes, Irreversibilities and Losses
An example of a simple reversible process is that shown in Figure 2.2, which depicts a ball
of mass m about to roll down a hill.
In position A, the ball has no kinetic energy, but a potential energy given by m × g × h,
where g is the acceleration due to gravity. If m is expressed in kg, g in m s−2 and h in m,
then the energy is expressed in joules.
In position B, the potential energy has been converted into kinetic energy. If there is no
rolling resistance or wind resistance, then the process is ‘reversible’, i.e., the ball can roll up
the other side and recover its potential energy.
In practice, however, some of the potential energy will be converted into heat because
of friction and wind resistance. The process is now ‘irreversible’ as the heat cannot be
converted back into kinetic or potential energy. It might be tempting to describe this as
a ‘loss’ of energy but that would not be very precise. In a sense, the potential energy is no
more ‘lost’ to heat than it is ‘lost’ to kinetic energy. So, the term ‘irreversible energy loss’ or
‘irreversibility’ is a rather more precise description of situations that many would describe
as a ‘loss of energy’.
Ball of mass m
A
h
B
Figure 2.2 Simple reversible process.
even exert a slight influence when no current is drawn, so the OCV of a fuel cell will
usually be lower than the value given by equation (2.9). Further explanation of ‘reversible’
and ‘irreversible’ processes is given in Box 2.2.
2.2
Open‐Circuit Voltage: Other Fuel Cells and Batteries
Equation (2.9) derived for the OCV of the hydrogen fuel cell is also applicable to other
reactions. The only step in the derivation that was specific to the hydrogen fuel cell
was the two electrons for each molecule of fuel consumed. In general terms, therefore,
equation (2.9) can be written as
Vr
g
zF
f
(2.11)
where z is the number of electrons transferred for each molecule of fuel.
The derivation is also not specific to fuel cells and applies equally well to other electrochemical power sources, particularly primary and secondary batteries. For example,
the primary alkaline cell that is used widely for domestic applications employs electrodes
of zinc and manganese dioxide. The overall cell reaction in this battery can be expressed
simply by
31
32
Fuel Cell Systems Explained
Zn 2MnO2 H2O
ZnO 2MnOOH
(2.12)
for which ∆ g f is −277 kJ mol−1.
At the negative electrode the reaction can be given as:
Zn 2OH
ZnO H2O 2e
(2.13)
and at the positive electrode as:
2MnO2 2H2O 2e
2MnOOH 2OH
(2.14)
Thus two electrons are passed round the circuit, and the OCV is expressed according
to equation (2.11), namely,
Vr
277 103
1.44 V
2 96 485
(2.15)
Another example is the methanol fuel cell, which is discussed in Chapter 6. The overall reaction is:
2CH3OH 3O2
4 H2O 2CO2
(2.16)
and involves the passage of 12 electrons from the negative to the positive electrode, i.e., 6
electrons for each molecule of methanol. For the methanol reaction, g f is −698.2 kJ mol−1.
Substituting this information into equation (2.11) gives:
Vr
698 103
1.21 V
6 96 485
(2.17)
It is to be noted that this is similar to the OCV for the hydrogen fuel cell.
2.3
Efficiency and Its Limits
The efficiency of a fuel cell—the fraction of the energy in the fuel that is converted into
useful electrical output—is a critical issue. Much is made of the fact that fuel cells are not
heat engines, so their efficiency is not limited by the Carnot cycle3 and therefore should be
high. This reasoning has driven much of the interest and investment in the technology.
The Carnot theorem as applied to a heat engine can be expressed as:
heat engine
W
H
T1 T2
T1
(2.18)
where W is the generated work, ΔH is the heat of combustion of the fuel and T1 and T2
are the absolute temperatures between which the heat engine operates. In practice, heat
engines are irreversible and normally operate with the lower temperature (T2) at room
temperature and with the upper temperature (T1) imposed by the materials of construction
of the engine. Thus, the efficiency of a heat engine is limited and depends on the
3 The Carnot cycle states that only a fraction of the heat produced by an engine can perform work and that
the remainder dissipates into the engine, its compartment and the environment.
Efficiency and Open‐Circuit Voltage
temperatures at which heat is supplied and withdrawn. As an example, for a steam turbine
operating at 400°C (673 K) with the water exhausted through a condenser at 50°C (323 K),
the Carnot efficiency limit is:
673 323
673
0.52 or 52%
(2.19)
For a fuel cell working ideally under isothermal conditions, the free energy change of
the reaction may be totally converted into electrical energy with a (maximum) efficiency
given by:
max
Wmax
H
G
H
1 T S
H
(2.20)
where Wmax is the maximum work delivered. The term TΔS is the heat exchanged with
the surroundings. Thus, under reversible conditions, the reaction enthalpy is converted
into electrical energy, except for an entropy term. The ΔH is usually larger in magnitude
than ΔG to such an extent that the ideal efficiency of a fuel cell, on a thermal basis, is
usually in the region of 90%, i.e., superior to that of a heat engine. It should be noted that,
for positive reaction entropies, the efficiency may become greater than 100% because
under isothermal conditions heat energy would be absorbed from the surroundings and
converted into electricity. The theoretical maximum efficiency of a fuel cell (ηmax) is
sometimes called the ‘thermodynamic efficiency’.
Unfortunately, the previously mentioned definition of efficiency is not without its
ambiguities, as there are two different values that can be used for the ΔH term. For the
conventional oxidation of hydrogen:
H2
hf
1
O
2 2
H2O(steam )
241.83 kJ mol
(2.21)
1
If the product water is condensed back to liquid, the reaction is:
H2
hf
1
O
2 2
H2O liquid
285.84 kJ mol
(2.22)
1
The difference between these two values for ∆h f (44.01 kJ mol−1) is the molar enthalpy
of vapourization4 of water. The higher figure is called the ‘higher heating value’ (HHV)
and the lower, quite logically, the ‘lower heating value’ (LHV).
Any statement of efficiency should say whether it relates to the HHV or LHV of the
fuel. When comparing the efficiencies of various appliances that are using the same fuel,
it is convenient to take the LHV since this is usually the maximum amount of heat that
can be recovered in the appliance itself. The difference between LHV and HHV varies
with the fuel. Generally, the sensible heat5 is small, and it is the heat of condensation of
steam that predominates. It follows that the richer the fossil fuel is in hydrogen, the
greater the deviation is between the LHV and the HHV. For example, the ratio of LHV
4 This used to be known as the ‘molar latent heat’.
5 Sensible heat is heat exchanged by a body or thermodynamic system in which the exchange of heat
changes the temperature of the body or system without causing a phase change.
33
Fuel Cell Systems Explained
Table 2.2 ∆ g f , maximum open‐circuit voltage and thermodynamic efficiency limit (HHV)
for hydrogen fuel cells.
Form of water
product
Liquid
Temperature (°C)
25
Liquid
∆g f
(kJ mol−1)
−237.2
Maximum open‐
circuit voltage (V)
Efficiency limit
(HHV) (%)
1.23
83
80
−228.2
1.18
80
Gas
100
−225.3
1.17
79
Gas
200
−220.4
1.14
77
Gas
400
−210.3
1.09
74
Gas
600
−199.6
1.04
70
Gas
800
−188.6
0.98
66
Gas
1000
−177.4
0.92
62
90
Fuel cell, liquid product
80
Efficiency limit/%
34
70
Fuel cell, steam product
60
Carnot limit, 50°C exhaust
50
40
30
0
200
400
600
800
1000
Operating temperature/°C
Figure 2.3 Maximum efficiency (HHV) of the hydrogen fuel cell at standard pressure. By way of
comparison, the Carnot limit is shown for a 50°C exhaust temperature.
to HHV is almost 1.0 for carbon monoxide (no hydrogen), 0.98 for coal (a little hydrogen),
0.91 for petrol, 0.90 for methane and 0.85 for hydrogen.
The values of the efficiency limit, relative to the HHV, for a hydrogen fuel cell are
listed in Table 2.2. The maximum OCVs, from equation (2.11), are also given. Plots in
Figure 2.3 show how efficiencies vary with temperature and how they compare with the
‘Carnot limit’. The following three important points should be noted:
1) Although the information displayed in Figure 2.3 and Table 2.2 would suggest that
lower fuel cell operating temperatures are better, the voltage losses are nearly always less
at higher temperatures (these losses are discussed in detail in Chapter 3). In practice,
therefore, fuel‐cell operating voltages are usually higher at higher temperatures.
Efficiency and Open‐Circuit Voltage
2) Any energy in the fuel that is not converted into electricity in the fuel cell appears as
waste heat (as with any heat engine). The waste heat from high‐temperature cells is
more useful than that from low‐temperature cells.
3) Contrary to statements often made by their supporters, fuel cells do not always have
a higher efficiency limit than heat engines.6
The decline in maximum possible efficiency with temperature associated with the
hydrogen fuel cell does not occur in exactly the same way with other types of fuel cell.
For example, when using carbon monoxide:
CO
1
O
2 2
CO2
(2.23)
The value of ∆ g changes even more rapidly with temperature, and the maximum possible efficiency falls from about 82% at 100°C to 52% at 1000°C. On the other hand, for
the reaction
CH 4 2O2
CO2 2H2O
(2.24)
∆ g is fairly constant with temperature, and therefore there is very little change in the
maximum possible efficiency.
Fuel‐cell efficiency is a topic that has given rise to much confusion in the literature. In
addition to the losses that originate in the cell stack, there are other system losses or
external inefficiencies to be taken into account. These include electrical losses in
compressing the incoming hydrogen and air and in converting the low‐voltage DC
output to high‐voltage AC. The total effect is a significant reduction in overall system
efficiency. Finally, if the fuel cells are to be used to propel electric vehicles, for example,
there are also inefficiencies in the electric motors and the drivetrain to be considered.
2.4
Efficiency and Voltage
It is clear from data given in Table 2.2 that there is a connection between the maximum
voltage of a cell and its maximum efficiency. The operating voltage of a fuel cell can also
be very easily related to its efficiency. This can be shown by adapting equation (2.9). If
all the energy from the hydrogen fuel, i.e., the heating value, or enthalpy of formation,
were transformed into electrical energy, the voltage would then be given by:
Vr
hf
2F
(2.25)
and have a value of 1.48 V and 1.35 V based on the HHV and the LHV, respectively.
These are the voltages that would be obtained for a 100% efficient system. Consequently,
the true efficiency of the cell is the actual voltage, Vc, divided by these values, e.g.,
Cell efficiency
Vc
100% HHV
1.48
(2.26)
6 In Chapter 8, it is shown how a heat engine and a high‐temperature fuel cell can be combined into a
particularly efficient system.
35
36
Fuel Cell Systems Explained
In practice, however, it is found that not all the fuel can be used, for reasons discussed
later; some of it usually has to pass through unreacted. A fuel utilization coefficient, μf,
can be defined as:
mass of fuel reacted in cell
f
(2.27)
mass of fuel input to cell
This parameter is equivalent to the ratio of the current delivered by the fuel cell to
that which would be obtained if all the fuel were reacted. The fuel‐cell efficiency, η, is
therefore given by:
f
Vc
100%
1.48
(2.28)
If a figure relative to the LHV is required, 1.25 instead of 1.48 should be used in the
previously mentioned formula. A good estimate for μf is 0.95, which allows the efficiency of a fuel cell to be estimated accurately from the very simple measurement of its
voltage. The efficiency can be a great deal less in some circumstances, as is discussed in
Section 2.5.3 and later in Chapter 6.
2.5
Influence of Pressure and Gas Concentration
2.5.1
Nernst Equation
As discussed in Section 2.1, the Gibbs free energy changes in a chemical reaction vary
with temperature. Equally important, though more complex, is the influence of both
reactant pressure and concentration. Consider, for example, a general reaction such as:
jA kB
mC
(2.29)
where j moles of A react with k moles of B to produce m moles of C. Each of the reactants,
as well as the product, has an associated ‘activity’,7 which is designated by the symbol a.
Accordingly, aA and aB represent the activities of the respective reactants, and aC the
activity of the product. For the case of gases behaving as ‘ideal gases’, it can be shown that
a
P
P
(2.30)
where P is the pressure, or partial pressure, of the gas and P° is the standard pressure,
namely, 100 kPa. Since fuel cells are generally gas reactors, this simple equation is
very useful. The activity of a gaseous component in the system can be taken to be
proportional to partial pressure, whereas for dissolved chemicals, the activity is linked
to the molarity (‘strength’) of the solution, which is usually expressed in mol dm−3. The case
7 The thermodynamic activity of a species is a measure of the ‘effective concentration’ of a species in a
reacting system. By convention, it is a dimensionless quantity. The activity of pure substances in condensed
phases (liquids or solids) is taken as unity. Activity depends principally on the temperature, pressure and
composition of the system. In reactions that involve real gases and mixtures, the effective partial pressure of
a constituent gas is usually referred to as ‘fugacity’.
Efficiency and Open‐Circuit Voltage
of the water produced in fuel cells is somewhat difficult, since this can be as either
steam or liquid. For steam, the following can be written:
PH2O
PH2O
aH2O
(2.31)
where PH2O is the vapour pressure of the steam at the temperature concerned; values for
this parameter can be obtained readily from published steam tables. When liquid water
is the product, it is a reasonable approximation to assume that aH2O 1.
The activities of the reactants and products modify the Gibbs free energy change of
a reaction. By using thermodynamic principles, for a chemical reaction such as the
general example given in equation (2.29), the following holds:
gf
gf
RT ln
aAj .aBk
aCm
(2.32)
where g f is the change in molar Gibbs free energy of formation at standard pressure.
For the reaction in a hydrogen fuel cell, equation (2.32) becomes:
1
gf
gf
RT ln
aH2 aO2 2
(2.33)
aH2O
The standard free energy change for the reaction ( g f ) is the quantity given in Tables 2.1
and 2.2. Thus, if the activity of the reactants increases, ∆ g f becomes more negative, i.e.,
more energy is released. On the other hand, if the activity of the product increases,
∆ g f increases and becomes less negative, and less energy is released. To see how activity
influences the cell voltage, ∆ g f can be substituted into equation (2.9) to obtain:
1
Vr
gf
2F
aH2 aO2 2
RT
ln
2F
aH2O
(2.34)
1
Vr
aH2 aO2 2
RT
ln
2F
aH2O
where Vr° is the OCV at STP. The equation shows precisely how raising the activity of
the reactants increases the voltage; it is known as the Nernst equation. Note that this
relationship is equally applicable to individual electrodes with potentials Er and E r°
replacing voltages Vr and Vr°, respectively.
The Nernst equation can be manipulated to investigate the influence of different
parameters on the operation and/or performance of a fuel cell. For example, in reaction
(2.21), namely,
H2
1
O
2 2
H2O steam
(2.21)
it can be assumed that the steam behaves as an ideal gas, and so:
aH2
PH2
, aO2
P
PO2
, aH2O
P
PH2O
P
(2.35)
37
38
Fuel Cell Systems Explained
Then the Nernst equation will become:
Vr
RT
ln
2F
Vr
PH2
PO2
P
P
1
2
(2.36)
PH2O
P
In nearly all cases, the pressures will be partial pressures; that is, the gases will be
components of a mixture. For example, the hydrogen gas might be part of a mixture of
hydrogen and carbon dioxide from a fuel reformer, together with product steam.
Oxygen will nearly always be a component of air. It is also often the case that the total
pressure on both the positive and negative electrodes is approximately the same as this
simplifies the cell design. If the system pressure is P, then
P , PO2
PH2
P , PHO
P
(2.37)
where α, β and δ are constants that depend on the molar masses and concentrations of
H2, O2 and H2O, respectively. The Nernst equation then becomes
Vr
Vr
RT
ln
2F
Vr
RT
ln
2F
1
2
1
(2.38)
P2
1
2
RT
ln P
4F
This relationship and equation (2.36) provide a theoretical basis for, and a quantitative indication of, the relative importance of a large number of variables in design and
operation of a fuel cell. These variables are discussed in more detail in later chapters,
but some points are considered briefly here to help introduce the technology.
2.5.2
Hydrogen Partial Pressure
Hydrogen can be supplied either pure or as part of a mixture. Isolation of the hydrogen
pressure term in equation (2.38) yields
1
Vr
Vr
P2
RT
ln O2
PH2O
2F
RT
ln PH2
2F
(2.39)
So, if the hydrogen partial pressure changes, say, from P1 to P2, with PO2 and PH2O
unchanged, then the resulting change in voltage ΔV will be given by
V
RT
ln P2
2F
P
RT
ln 2
2F
P1
RT
ln P1
2F
(2.40)
The use of hydrogen mixed with carbon dioxide occurs particularly in phosphoric
acid fuel cells (PAFCs) that operate at about 200°C (473 K). Substituting the values for
R, T and F in equation (2.40) yields:
Efficiency and Open‐Circuit Voltage
V
P2
P1
0.02 ln
(2.41)
This relationship gives values that are in good agreement with experimental results,
which correlate best with a factor of 0.024 instead of 0.020. As an example, changing
from pure hydrogen to a 50% H2–50% CO2 mixture causes a reduction of 0.015 V per cell.
2.5.3
Fuel and Oxidant Utilization
As air passes through the positive‐electrode (cathode) compartment of a fuel cell, oxygen is consumed, and thereby its partial pressure is reduced. Similarly, the partial pressure of the fuel will often decline in the negative‐electrode compartment. Referring to
equation (2.39), it can be seen that α and β decrease, whereas δ increases. Consequently,
the following term in equation (2.38):
RT
ln
2F
1
2
(2.42)
becomes smaller as fuel and oxidant are consumed as they pass through the cell, and so
the cell voltage would be expected to fall between the inlet and the outlet of the cell. In
most stack designs, it is not actually possible to have variations in voltage throughout
a cell—the fact that the electrodes are good electronic conductors ensures that the
voltage is approximately uniform throughout each cell. Accordingly, it is the current
density that changes throughout the cell. The current density will be lowest nearer the
exit where the fuel concentration is lower.8
The RT term in equation (2.42) also dictates that the drop in cell voltage (or current
density where the voltage cannot change) due to fuel and oxidant being consumed will
be greater in high‐temperature fuel cells.
Obviously, for a system to exhibit high efficiency, the fuel utilization should be as high
as possible. On the other hand, equation (2.39) also suggests that high fuel utilization
will lead to low average cell voltage or current density. The effect of low current density
can be compensated by increasing the size of the cell, but this will increase the cost. In
a practical system, therefore, it is always necessary to reach a compromise between fuel
utilization and stack size (i.e., cost). This issue is most important with high‐temperature
cells and is considered further in Chapters 7–9.
2.5.4
System Pressure
The Nernst equation also demonstrates that the system pressure can increase the
voltage of a fuel cell according to the term:
RT
ln P
4F
(2.43)
8 The current density distribution in a stack will depend also on the orientation of the fuel and oxidant
channels. Where the flows are parallel and in the same direction (co‐flow), the current density will be lowest
at the outlet of the cells. This is not the case for counter‐flow or cross‐flow configurations. Modern flow‐
field design is focused on optimizing current density distribution throughout the stack.
39
40
Fuel Cell Systems Explained
For instance, if the pressure changes from P1 to P2, there will be change in voltage:
V
RT
P
ln 2
4F
P1
(2.44)
For a solid oxide fuel cell operating at 1000°C, the equation would give:
V
0.027 ln
P2
P1
(2.45)
This relationship has been found to be in very good agreement with reported results for
high‐temperature cells, but not for other fuel cells that work at lower temperatures. For
example, whereas a PAFC at 200°C should be affected by system pressure according to:
V
0.010 ln
P2
P1
(2.46)
published data deliver a different correlation, namely,
V
0.063 ln
P2
P1
(2.47)
In other words, at lower temperatures, the benefits of raising system pressure are much
greater than predicted by the Nernst equation. This discrepancy in performance is
because, except for very high‐temperature cells, increasing the pressure also reduces
the losses at the electrodes, especially at the positive electrode; see Chapter 3.
A similar outcome occurs when changing the oxidant from air to oxygen. This action
effectively changes β in equation (2.38) from a value of 0.21 (21% oxygen in air) to 1.0
(pure oxygen). Isolating β in this equation gives:
Vr
Vr
RT
ln
4F
RT
ln
2F
RT
ln P
4F
(2.48)
The change in β from 0.21 to 1.0, with all other factors remaining constant, yields:
V
RT
1.0
ln
4F
0.21
(2.49)
For a PEMFC at 80°C, the change in voltage would be 0.012 V. In fact, studies have
demonstrated a much larger change; namely, 0.05 V is commonplace. Again, this is due to
the reduction in overpotential at the cathode (positive electrode) as a result of high oxygen
pressure.
2.6
Summary
The OCV (also known as the reversible voltage) for a hydrogen fuel cell is given by:
Vr
G
2F
(2.50)
Efficiency and Open‐Circuit Voltage
where ΔG is the free energy change for the fuel‐cell reaction. In general, for a reaction
where z electrons are transferred for each molecule of fuel, the OCV is:
Vr
G
zF
(2.51)
The Gibbs free energy change, ΔG, varies with temperature and other factors. The
maximum efficiency is given by the expression:
G
100%
H
max
(2.52)
The efficiency (HHV) of a working hydrogen fuel cell can be found by using the following simple formula:
f
Vc
100%
1.48
(2.53)
where μf is the fuel utilization (typically about 0.95) and Vc is the voltage of a single cell.
The pressure and concentration of the reactants also influence the change in Gibbs
free energy, and thus the voltage. This is expressed in the Nernst equation, which can
take many forms. For example, if the water product is in the form of steam, then:
1
Vr
Vr
PH2 PO2 2
RT
ln
2F
PH2O
(2.54)
where Vr° is the cell OCV at standard pressure.
In most of this chapter, equations have been given for the voltage of a cell, or its OCV.
In practice the operating voltage is less than that predicted and in some cases much less.
This is the result of losses or ‘irreversibilities’, which are explained more fully in the next
chapter.
Further Reading
Barclay, FJ, 2006, Fuel Cells, Engines and Hydrogen: An Exergy Approach, John Wiley &
Sons, Ltd, Chichester. ISBN: 978‐0‐470‐01904‐7.
EG&G Technical Services, Inc., under contract to US Department of Energy, 2016, Fuel Cell
Handbook (Seventh Edition), National Energy Technology Laboratory, Morgantown, WV.
Srinivasan, S, 2006, Fuel Cells. From Fundamentals to Applications, Springer, New York.
ISBN: 9781441937728.
Stolten, D (ed.), 2010, Hydrogen and Fuel Cells – Fundamentals, Technologies and
Applications, Wiley‐VCH, Verlag GmbH & Co. KGaA, Weinheim. ISBN: 978‐3‐527‐32711.
41
43
3
Operational Fuel‐Cell Voltages
3.1
Fundamental Voltage: Current Behaviour
As shown in Chapter 2, the theoretical value of the ‘no-loss’ open‐circuit voltage of a
hydrogen fuel cell is expressed by equation (2.9):
Vr
gf
2F
(2.9)
where ∆ g f is the change in free energy for the cell reaction (i.e., the difference between
the free energy of formation of the reactants and the free energy of formation of the
products) and F is the Faraday constant. This gives a value of about 1.2 V for a cell that
is operating below 100°C. When, however, a fuel cell is put to use, it is found that the
‘operational voltage’ is less than this, indeed often considerably less. The voltage versus
current density1 performance of a single cell of typical design and operating at 40°C and
normal air pressure is presented in Figure 3.1. The key points are as follows:
●
●
●
●
Even the open‐circuit voltage is less than the theoretical value.
There is a rapid initial drop in voltage.
The voltage then falls less slowly and more linearly.
A more rapid decline in voltage may be observed at higher current densities.
There are two marked changes in the previously mentioned performance characteristics when a fuel cell is operated at higher temperatures, namely:
●
●
As shown in Chapter 2, the reversible (‘no loss’) voltage falls, and thereby its value
usually becomes closer to that of the actual operating voltage.
The initial drop in voltage as current is drawn from the cell is greatly reduced.
The performance for a typical solid oxide fuel cell (SOFC) that is operating at about
800°C is given in Figure 3.2 and has the following significant features:
●
●
●
The open‐circuit voltage is equal to, or only a very little less than, the theoretical value.
The initial drop in voltage is very small, and the graph is considerably more linear.
There may be a higher current density at which the voltage falls rapidly away, as found
for fuel cells that run at lower temperatures.
1 It is common practice to refer to current density, or current per unit area, rather than just current so that it
is easier to compare the performance of cells of different size.
Fuel Cell Systems Explained, Third Edition. Andrew L. Dicks and David A. J. Rand.
© 2018 John Wiley & Sons Ltd. Published 2018 by John Wiley & Sons Ltd.
Fuel Cell Systems Explained
‘No-loss’ voltage of 1.2 V
1.2
Even the open-circuit voltage is less than the
theoretical no-loss value
1.0
Rapid initial fall in voltage
Cell voltage/V
44
Voltage falls more slowly,
and graph is fairly linear
0.8
0.6
0.4
Voltage begins to fall faster
at higher currents
0.2
0
0
200
400
600
800
Current density/mA cm–2
1000
Figure 3.1 Voltage versus current density performance of a typical fuel cell operating at low
temperature and air pressure.
Comparison of the two sets of data reveals that although the reversible voltage is lower
for the cell running at the higher temperature, the actual operating voltage is generally
greater, because the voltage drop or ‘irreversibilities’ are smaller.
This chapter examines the factors that are responsible for the voltage falling below the
reversible value and consider ways to ameliorate their adverse effects.
3.2 Terminology
Efforts to develop fuel‐cell systems are highly interdisciplinary. Success requires the
skills of chemists, electrochemists, materials scientists, thermodynamicists, electrical
and chemical engineers, control and instrumentation engineers and others. Not
surprisingly, there are occasions when these various disciplines have their own names
for what is often essentially the same performance parameter. The main topic of this
chapter — fuel‐cell voltage — is a case in point.
The graphs of Figures 3.1 and 3.2 show the difference between the voltage that is
expected from a fuel cell operating reversibly (ideally) and the voltage that is
observed in practice. Remarkably, five names are commonly used to denote the
voltage difference:
●
‘Overvoltage’ is a term often adopted by electrochemists to describe the nonideal
behaviour of electrolysers, fuel cells and batteries. Similarly, ‘overpotential’ signifies
differences in potentials that are generated at electrode|electrolyte interfaces.
Unfortunately, the form of the word overvoltage tends to imply that the observed
voltage is larger than the value predicted by theory, whereas in fuel cells the observed
voltage is smaller.
Operational Fuel‐Cell Voltages
1.2
‘No-loss’ voltage of 1.0 V
1.0
Cell voltage/V
Graph is fairly linear
0.8
0.6
Very small initial fall in voltage, and open-circuit
voltage is very close to theoretical value
0.4
Voltage begins to fall faster
at higher currents
0.2
0
0
200
400
600
800
1000
Current density/mA cm–2
Figure 3.2 Voltage versus current density performance of a typical fuel cell operating at about 800°C
and air pressure.
●
●
●
●
‘Polarization’ is another term that has been employed by electrochemists, but it is
misleading on several counts and is generally best avoided.
‘Irreversibility’ is the best term from a thermodynamics point of view. Nonetheless, it
is perhaps not sufficiently specific to fuel cells and does not connect well with the
main effect under consideration here, namely, that which gives rise to a reduction in
cell voltage.
‘Voltage loss’ may be taken as a simple way to indicate that a practical fuel cell
exhibits a voltage that is less than would be expected from thermodynamic
considerations. A discussion of ‘reversibility, irreversibility and losses’ is given in
Section 2.1, Chapter 2.
‘Voltage drop’ is certainly not scientifically precise, but it does convey the effect
observed and is readily understood by electrical engineers.
These alternative terms, which demonstrate the richness of the English language
often having many words for the same subject, will be encountered during the course of
this book.
It is also worth remarking that ‘potential’ and ‘voltage’ are often misleadingly used
interchangeably. As with the thermodynamic properties G, H and S that were introduced
in the last chapter, electric potential is only able to be measured as a potential difference
between two electrodes. Since a standard state is defined for thermodynamic properties,
electrochemists have adopted the standard hydrogen electrode (SHE) as the reference
against which the potential of an electrode can be measured. In this book, E is used to
denote the potential of an electrode (i.e., with reference to the SHE), and E° is the
electrode potential under standard conditions. The voltage difference between two
electrodes in a cell is represented by the symbol V.
45
46
Fuel Cell Systems Explained
3.3
Fuel‐Cell Irreversibilities
The characteristic shape of the voltage versus current density relationships shown in
Figures 3.1 and 3.2 is the result of four major irreversibilities. The concomitant voltage
loses will be outlined briefly here before being considered in more detail later, namely:
1) Activation losses. These represent the slowness of the reactions taking place on the
surface of the electrodes. A proportion of the voltage generated is lost in driving the
chemical reaction that transfers the electrons to or from the electrode. As discussed
in Section 3.4, the resulting effect on the voltage is highly non‐linear.
2) Internal currents and fuel crossover. This voltage loss results from a small amount of
fuel passing through the electrolyte from the anode to the cathode and, to a lesser
extent, from electron conduction through the electrolyte. In an ideal situation, the
electrolyte should only transport ions through the cell, as illustrated in Figures 1.3
and 1.4, Chapter 1. In practice, however, a certain amount of fuel diffusion and
electron flow will always be possible. Generally, the fuel loss and current are both
small, and thereby the net effect is usually not very important. Crossover does,
however, have a marked influence on the open‐circuit voltage of low‐temperature
cells, as will be examined in Section 3.5.
3) Ohmic losses. This voltage loss is the straightforward resistance to the flow of
electrons through the material of the electrodes and the various interconnections, as
well as the resistance to the flow of ions through the electrolyte. The voltage drop is
essentially linearly proportional to the current density and therefore is sometimes
also called resistive losses.
4) Concentration or mass‐transport losses. These losses arise from the change in
concentration of the reactants at the surface of the electrodes as the fuel is
consumed. Since reactant concentration affects the voltage, this type of irreversibility is sometimes referred to as concentration losses. Because the effect is the
result of a failure to transport sufficient reactant to the electrode surface, the
term mass‐transport losses is also used. There is even a third name — Nernstian
losses — that evolved following modelling of the effects of concentration by the
Nernst equation.
The four categories of irreversibility are considered, in turn, in the sections that follow.
3.4
Activation Losses
3.4.1 The Tafel Equation
In considering the overvoltage at any one electrode, the activation loss (ΔEact) can be
defined as:
E act
E Eeq
(3.1)
where E is the measured electrode potential and Eeq is the theoretical equilibrium
electrode potential. As a result of experiments rather than theoretical considerations,
Julius Tafel observed and reported in 1905 that the variation in potential (later to be given
Operational Fuel‐Cell Voltages
Equation of best-fit line is
V = a log (i/io)
0.6
0.5
Overpotential/V
0.4
Fast reaction
Slow reaction
0.3
0.2
0.1
0
0
1
2
3
4
5
Log (current density)/mA cm–2
Best-fit line intercepts the current density axis at io
Figure 3.3 Tafel plots for slow and fast electrochemical reactions.
the term overpotential2) at the surface of an electrode followed a similar pattern for a
great variety of electrochemical reactions.
This general behaviour, which is displayed in Figure 3.3, shows that if overpotential is
plotted against the log of current density, then, for most values of overpotential, the
relationship approximates to a straight line. Such a graph is known as a ‘Tafel plot’ and
the linear relationship is represented by the expression:
E act
a log
i
io
(3.2)
where a is a constant, commonly referred to as the ‘Tafel slope’, i is the current density
and io is the ‘exchange‐current density’, i.e., the current density at zero overpotential or
that at which the overpotential begins to manifest itself.
The exchange‐current density io can be visualized as follows. The reaction at the
oxygen electrode of a proton‐exchange membrane or acid electrolyte fuel cell is:
O2 4 e
4H
2H 2 O
(3.3)
At zero current density, it may be assumed that there is no activity at the electrode,
and therefore this reaction does not take place. In fact, this is not so. The reaction is
2 Agar, JN and Bowden, FP, 1938, The kinetics of electrode reactions I and II, Proceedings of the Royal
Society of London. Series A, Mathematical and Physical Sciences, vol. 169 (937), pp. 206–234.
47
48
Fuel Cell Systems Explained
occurring, but the reverse reaction is also proceeding at the same rate. There is an
equilibrium, which is expressed as:
4 H 2H 2 O
O2 4 e
(3.4)
Thus, there is a continual backwards and forwards flow of electrons from and to the
electrolyte that constitutes the exchange‐current density, io. If the value of io is high,
then the surface of the electrode can be said to be more ‘active’, leading to a low
activation loss when current is drawn. If the value of io is low, the activation overpotential
will be high.
Equation (3.2) is known as the Tafel equation and can be expressed in many forms.
One simple and preferred variation is to use natural logarithms instead of base‐10
logarithms, i.e.,
Eact
A ln
i
io
(3.5)
The constant A is higher for an electrochemical reaction that is slow. It is important to
remember that the Tafel equation only holds true when i > io.
3.4.2 The Constants in the Tafel Equation
Although it was originally deduced from experimental results, the Tafel equation also has
a theoretical basis. For a hydrogen fuel cell, the constant A in equation (3.5) is given by:
A
RT
2 F
(3.6)
where R is the universal gas constant (= 8.314 472 J K−1 mol−1) and T is the temperature
in Kelvin (K). The parameter α is called the ‘charge‐transfer coefficient’ and is the
proportion of the electrical energy applied that is harnessed in changing the rate of an
electrochemical reaction. Its value depends on the reaction involved and the material
used for the electrode, but it must be in the range 0–1.0. For the hydrogen electrode, α
is about 0.5 for a wide variety of electrode materials. At the oxygen electrode, the
charge‐transfer coefficient shows more variation but is still between about 0.1 and 0.5
in most circumstances. In short, experimenting with different materials to get the best
possible value for A will make little impact.
The presence of T in equation (3.6) might give the impression that raising the temperature increases the overpotential. In fact, this is very rarely the case as the effect of
increases in io with temperature far outweighs any increase in A. Indeed, the key to
making the activation overpotential as low as possible is the value of io, as this can vary
by several orders of magnitude. Furthermore, io is influenced by several parameters
other than the material used for the electrode. In summary, the exchange‐current
density is crucial in controlling the performance of a fuel‐cell electrode.
Equations (3.5) and (3.6) can be rearranged to describe the cell current as a function
of voltage. This is achieved by converting from the logarithmic to the exponential
form, to give:
i io exp
2 F Eact
RT
(3.7)
Operational Fuel‐Cell Voltages
Electrochemists will recognize this as a form of the Butler–Volmer equation that is
more fully expressed as:
i io exp
n
aF
RT
Eact
exp
n c F Eact
RT
(3.8)
where n is the number of electrons transferred in the electrochemical reaction and
αa and αc are the charge‐transfer coefficients at the negative and positive electrodes,
respectively. The Butler–Volmer equation is one of the most fundamental equations in
electrochemistry as it expresses the current produced by an electrochemical reaction in
terms of the rates of reactions at the two electrodes. The equation was derived from
kinetic theory and provides a sound basis for its simpler but empirical relative — the
Tafel equation, which only holds when the exchange‐current density is very much
smaller than the measured current density (i >> io). Even so, the Tafel equation is
adequate for understanding and expressing the performance of most practical fuel‐cell
systems.
For a fuel cell that has no losses except for the activation overpotential on one electrode, the cell voltage would be given by:
Vc
Vr
A ln
i
io
(3.9)
where Vr is the open‐circuit voltage given by equation (2.9). Plots of cell voltage (Vc)
versus current density (i) obtained using equation (3.9) with values of io of 0.01, 1.0
and 100 mA cm−1 and using a typical value for A of 0.06 V are presented in Figure 3.4.
‘No-loss’ voltage of 1.2 V
1.2
100
1.0
Cell voltage/V
0.8
1.0
0.6
0.01
0.4
0.2
0
0
200
400
600
800
1000
Current density/mA cm–2
Figure 3.4 Cell voltage versus current density, assuming losses due only to the activation
overpotential at one electrode, for exchange‐current density io values of 0.01, 1.0 and 100 mA cm–2.
49
50
Fuel Cell Systems Explained
The importance of io can be clearly seen. High values of io give the highest actual cell
voltages, and low values result in the lowest cell voltages. For most values of current
density, the actual cell voltage is fairly constant for each value of io. Note that when io
is 100 mA cm−2, there is no voltage drop until the current density i is greater than
100 mA cm−2.
It is possible to measure the overpotential at each electrode, either with reference
electrodes within a working fuel cell, or by using half cells, as described later. The values
of io for the hydrogen electrode at 25°C for various metal substrates are given in Table 3.1;
the measurements were conducted on flat smooth electrodes. The great variation in
exchange-current indicates that some metals are more catalytically active than others.
There is often inconsistency between values obtained by different researchers, which
suggests that there are several influencing factors. The io for the cathode also varies
appreciably and is generally lower than that for the anode by a factor of about 105. For a
cathode, therefore, the exchange-current is of the order of 10−8 A cm−2, even when using
a platinum catalyst, i.e., far lower than the lowest curve in Figure 3.4. Fortunately, in
practice, the value of io for a fuel‐cell electrode is much higher than those given in
Table 3.1 because the roughness of the electrode makes the ‘real’ surface area many times
larger (typically, by at least three orders of magnitude) than the nominal length × width.
The differences in values of io between the two electrodes reflect the different rates of
the reactions that take place on either side of the cell. The hydrogen oxidation reaction
(HOR) on the anode is a very fast and simple reaction. By contrast, the oxygen reduction
reaction (ORR) on the cathode is many times slower because it is more complex, i.e.,
several reaction steps are involved. It is generally considered that the overpotential at
the anode is negligible compared with that at the cathode, at least in the case of hydrogen
fuel cells.
Table 3.1 Values of io for the hydrogen electrode
for various metals in an acid electrolyte.
Metal
io (A cm−2)
Pb
2.5 × 10−13
Hg
3 × 10−12
Zn
3 × 10−11
Cd
8 × 10−10
Mn
1 × 10−11
Ti
2 × 10−8
Ta
1 × 10−7
Mo
1 × 10−7
Fe
1 × 10−6
Ag
4 × 10−7
Ni
6 × 10−6
Pt
5 × 10−4
Pd
4 × 10−3
Operational Fuel‐Cell Voltages
In other fuel cells, for example, the direct methanol fuel cell (DMFC), the overpotential at the anode is by no means negligible. In these systems, the equation for
the total activation overvoltage would combine contributions from both electrode
polarities, namely,
Activation overvoltage
Aa ln
i
ioa
Ac ln
i
(3.10)
ioc
where ioa and ioc are the exchange‐current densities at the anode and cathode, respectively.
This equation can be expressed as:
V
A ln
i
b
(3.11)
where ΔV is the total drop in voltage due to the combined activation overpotentials and
A
Aa
Ac
Aa
Ac
and b ioaA
iocA
(3.12)
Note that the equation (3.12) is only valid for i > b. The relationship mimics equation
(3.5), which expresses the overpotential for one electrode. So whether the activation
overpotential arises mainly at one electrode only, or both, the equation that models the
voltage is of a similar form. Moreover, in all cases, the term in the equation that shows the
most variation is the exchange‐current density io, rather than the parameter A. Further
discussion of electrode kinetics for different fuel‐cell types will appear in later chapters.
3.4.3
Reducing the Activation Overpotential
Improving fuel‐cell performance via increasing the value of io can be accomplished in
various ways:
●
●
●
●
●
Raising the cell temperature. This action fully explains the different shape of the
voltage versus current density graphs of low‐ and high‐temperature fuel cells
illustrated in Figures 3.1 and 3.2. For a low‐temperature cell, the io at the positive
electrode will be about 0.1 mA cm−2, whereas for a typical 800°C cell, it will be about
10 mA cm−2 — a 100‐fold improvement!
Using more effective catalysts. The effect of different metals in the electrode is
shown clearly by the data given in Table 3.1 where the precious metals platinum and
palladium are much more active for hydrogen activation than base metals such as
zinc and lead. In recent years, major efforts have been made to develop superior
catalysts through the use of alloys.
Increasing the roughness of the electrodes. This technique increases the real surface
area of each nominal 1 cm2 that, in turn, enhances the io.
Increasing reactant concentration, e.g., using pure oxygen instead of air. Such
action enables the catalyst sites to be more effectively occupied by reactants. As
demonstrated in Chapter 2, this also increases the open‐circuit voltage.
Increasing the pressure. This approach is also considered to be effective through
enhancing the reactant occupancy of catalyst sites. Similar to enhancing the reactant
concentration, the strategy produces a ‘double benefit’ through increasing the open‐
circuit voltage.
51
52
Fuel Cell Systems Explained
The last two points in this list explain the discrepancy between the theoretical and the
actual open‐circuit voltage that has been discussed in Section 2.5.4, Chapter 2.
It is useful to reflect that the activity of the catalyst, the electrode roughness and the
issues of pressure and reactant concentration all exert an influence on the reaction rate
and, consequently, on the performance of the fuel cell. The electrode reactions take
place at a triple‐phase boundary, and therefore cell performance is highly dependent on
the design and distribution of the catalyst and its interaction with the electrode (i.e., the
catalyst topology). Greater consideration will be given to such requirements when
examining each type of fuel cell in later chapters and will include the introduction of
advanced solid‐state materials such as mixed ionic–electronic conductors.
3.5
Internal Currents and Fuel Crossover
Although the electrolyte of a fuel cell will have been chosen for its ion‐conducting properties, it will invariably possess some electronic conductivity. Minute internal currents due
to conduction of electrons will reduce the cell voltage by a small amount. Probably more
important in a practical fuel cell is that some hydrogen will diffuse from the anode, through
the electrolyte, to the cathode. The hydrogen will react directly with oxygen on the
cathode catalyst to be consumed and thereby generate no current from the cell. The
wasted fuel that migrates in this manner through the electrolyte is known as ‘fuel
crossover’.
The previously mentioned two adverse effects are essentially equivalent. The crossing
over of one hydrogen molecule wastes two electrons and amounts to exactly the same
as two electrons crossing internally in the opposite direction rather than as an external
current. Furthermore, if the major loss in the cell is the transfer of electrons at the
interface of the cathode, which is the case for hydrogen fuel cells, then the effect of
both these phenomena on the cell voltage is also the same.
Internal electron flow or fuel crossover will typically be the equivalent of only a few
mA cm−2. In terms of energy loss, the irreversibility is not very important. In low‐
temperature cells, however, it does cause a very noticeable voltage drop under open‐
circuit conditions. Users of fuel cells can readily accept that the working voltage of a cell
will be less than the theoretical ‘no loss’ reversible voltage. In an open circuit, however,
when no work is being done, it may be expected that the cell voltage will be the same as
the reversible voltage. For low‐temperature cells, such as proton‐exchange membrane
fuel cells (PEMFCs), when operating on air at ambient pressure, the open‐circuit
voltage will usually be at least 0.3 V less than the reversible voltage (~1.2 V), due to
internal currents or crossover.
If, as in the last section, the losses in a fuel cell are assumed to be caused only by the
‘activation overpotential’ at the cathode, then the cell voltage (Vc) will be reduced only
by the amount given by equation (3.9), namely:
Vc
Vr
A ln
i
io
(3.9)
For a PEMFC operating at about 30°C and using air at atmospheric pressure, reasonable
values for the parameters in equation (3.9) are V = 1.2 V, A = 0.06 V and io = 0.04 mA cm−2.
Operational Fuel‐Cell Voltages
Table 3.2 PEMFC voltages at low current densities.
Current density (mA cm−2)
Voltage (V)
0
1.2
0.25
1.05
0.5
1.01
1.0
0.97
2.0
0.92
3.0
0.90
4.0
0.88
5.0
0.87
6.0
0.86
7.0
0.85
8.0
0.84
9.0
0.83
If the internal current density
is 1.0 mA cm–2, then the
open-circuit voltage will drop
to 0.97 V
Using these values, the cell voltages for a range of low current densities have been
calculated and are listed in Table 3.2.
Because of the internal currents, the current density is not zero, even if the cell is at
open circuit. For instance, if the internal current density is 2 mA cm−2, then the open‐
circuit voltage would be 0.92 V, i.e., nearly 0.3 V (or 25%) less than the theoretical value.
This appreciable loss in voltage is a consequence of the very steep initial fall that is
shown by the data in Figure 3.4 (v.s.). The steepness of the curve also explains why the
open‐circuit voltage of low‐temperature fuel cells is highly variable. The information
given in Table 3.2 and Figure 3.4 demonstrates that a small change in fuel crossover and/
or internal current caused, for example, by a change in the humidity of the electrolyte,
can promote a large change in open‐circuit voltage.
Obviously, it is not easy to measure the fuel crossover and the internal current — an
ammeter cannot be inserted in the circuit! One method, however, is to determine the
consumption of reactant gases at open circuit. For single cells and small stacks, the very
low rates of gas usage cannot be measured by means of normal gas flow meters, so that
bubble counting, gas syringes, or similar have to be employed. For example, at open
circuit, a small PEM cell of area 10 cm2 might have a hydrogen consumption of
0.0034 cm3 s−1, at normal temperature and pressure (author’s measurement performed
on a commercial cell). According to Avogadro’s law, the volume of 1 mol of any gas is
2.24 × 104 cm3 at standard temperature and pressure (STP), and therefore the gas usage
is 1.52 × 10−7 mol s−1. Equation (A2.13) in Appendix 2 shows that the rate of hydrogen
fuel usage in a single cell (n = 1) is related to the current (I) by the formula:
Gas usage
I
mol s
2F
1
(3.13)
The previously mentioned losses therefore correspond to a current of 1.52 × 10−7 × 2 ×
9.65 × 104 = 29 mA. Given that the cell area is 10 cm2, the current density is 2.9 mA cm−2
53
Fuel Cell Systems Explained
and is the sum of the current equivalent of fuel lost from crossover and the actual internal
current density. If in is the value of this internal current density, then equation (3.9) used
to express the cell voltage can be refined to:
Vc
Vr
A ln
i in
io
(3.14)
Taking typical values for a low‐temperature cell, namely, V = 1.2 V, A = 0.06 V,
io = 0.04 mA cm−2 and in = 3 mA cm−2, yields a graph of cell voltage against current
density of the form displayed in Figure 3.5; the relationship is quite similar to that shown
in Figure 3.4. The importance of the internal current is considerably less for high‐
temperature cells because the exchange‐current density io is very much greater and,
consequently, the initial fall in voltage is less dramatic.
3.6
Ohmic Losses
The losses in cell voltage due to the electrical resistance of the electrodes, and to the
resistance to the flow of ions in the electrolyte, are the simplest to understand and to
model. The size of the voltage drop (ΔV) is simply proportional to current, i.e., as given
by Ohm’s law:
V
(3.15)
IR
In most fuel cells, the resistance — R in equation (3.15) — mainly emanates from the
electrolyte, though the cell interconnects or bipolar plates (see Section 1.3, Chapter 1)
can also be important contributors.
‘No-loss’ voltage of 1.2 V
1.2
1.0
0.8
Cell voltage / V
54
0.6
0.4
0.2
0
0
200
400
600
Current density/mA
800
1000
cm–2
Figure 3.5 Fuel‐cell voltage modelled using activation and fuel crossover/internal current losses only.
Operational Fuel‐Cell Voltages
To be consistent with the other equations for voltage loss, equation (3.15) should be
expressed in terms of current density. To do this, it is necessary to introduce the
concept of the resistance corresponding to 1 cm2 of the cell. The parameter is called the
‘area specific resistance’ (ASR) and can be represented by the symbol r. The equation
for the voltage drop then becomes:
(3.16)
V ir
where i is, as usual, the current density in mA cm−2 and therefore r should be given in
kΩ cm2.
Using the methods described in Section 3.10, it is possible to distinguish this particular irreversibility from the others. For instance, it can be shown that the ‘ohmic loss’ of
voltage is significant in all types of cell and is especially important in the case of the
SOFC. Three ways of reducing the internal resistance of a cell are as follows:
●
●
●
The use of electrodes with the highest possible conductivity.
Optimization of the design and choice of materials for the bipolar plates or cell interconnects. This issue has already been addressed in Section 1.3, Chapter 1.
Making the electrolyte as thin as possible. Unfortunately, such an approach is often
difficult given that if a solid electrolyte is employed, it sometimes has to be fairly thick
as it is the support on which the electrodes are built. Also where the electrolyte is a
liquid, e.g., in the alkaline fuel cell, the separation of electrodes has to be sufficiently
wide to allow a circulating flow of electrolyte between them. The electrolyte in a
SOFC can be made very thin but still must have adequate thickness to prevent
internal shorting between electrodes, a requirement that implies a certain level of
physical robustness.
3.7
Mass‐Transport Losses
If the oxygen at the positive electrode of a fuel cell is supplied in the form of air, then
it is self‐evident that during operation, there will be a slight reduction in the concentration of the oxygen in the region of the electrode, as the reactant gas is extracted.
The extent of the change in concentration, which reduces the partial pressure of
oxygen, will depend on the current being taken from the fuel cell and on physical
factors that relate to how well the air around the electrode can circulate and how
quickly the oxygen can be replenished. Similarly, if the negative fuel electrode is
supplied with a gas mixture that contains hydrogen (such as a reformed gas containing
carbon oxides), then there will be a fall in hydrogen partial pressure as the hydrogen
is consumed by the cell. Whether addressing a reduction in the absolute pressure or
in the partial pressure, the same principles apply, and the net result will be a reduction
in voltage.
There is no analytical solution to modelling the change in cell voltage as a function of
the hydrogen partial pressure. One approach is to revisit the Nernst equation, i.e., use
equation (2.40) in Chapter 2:
V
RT
P
ln 2
2F
P1
(2.40)
55
56
Fuel Cell Systems Explained
Note that this particular equation relates the increase in cell voltage due to increasing
pressure from P1 to P2. The equation can be used to estimate the voltage drop as a
result of the decrease in pressure caused by consumption of the fuel gas as follows.
Consider a limiting value of current density, il, at which fuel is consumed at a rate
equally to its maximum supply rate. Clearly, the current density cannot rise above this
value because the fuel gas cannot be supplied at a greater rate. At this current density,
the pressure of the hydrogen supply will have just fallen to zero. If P1 is the pressure
when the current density is zero, and it is assumed that the pressure falls linearly
down to zero at the current density il, and then the pressure P2 at any current density
i is given by:
P2
P1 1
i
il
(3.17)
Substitution of this relationship into equation (2.40), given earlier, yields the voltage
change due to the concentration (mass‐transport) losses, namely:
V
RT
i
ln 1
il
2F
(3.18)
It should be noted that care must be taken over the signs, i.e., equations (2.40) and (3.18)
are written in terms of a voltage gain, and the term inside the brackets is always less
than 1. Consequently, the equation for voltage drop should be written as:
V
i
RT
ln 1
il
2F
(3.19)
More generally, the concentration (mass‐transport) losses are given by:
V
B ln 1
i
il
(3.20)
where B is a parameter that depends on the fuel cell and its operating state.
For example, if B is set to 0.05 V and il to 1000 mA cm−2, then quite a good fit is obtained
to curves such as those in Figures 3.1 and 3.2. Nevertheless, this theoretical approach
has many weaknesses, especially in the case of the vast majority of fuel cells, which are
supplied with air rather than oxygen.
An alternative way to quantify the voltage loss is to use an empirical relationship, e.g.,
V
m exp ni
(3.21)
where m and n are constants. Using values of m = 3 × 10−5 V and n = 8 × 10−3 cm2 mA−1, the
voltage change predicted by equations (3.20) and (3.21) are very similar. In particular,
equation (3.21) is found to give a good fit with voltage losses that are measured
experimentally and is widely accepted in the fuel‐cell community. It will be used in the
sections that follow.
The overvoltage due to concentration (mass‐transport) losses is particularly important
in cases where the hydrogen is supplied from a reformer or generator; as such an
arrangement might have difficulty in adjusting the rate of supply of hydrogen sufficiently
rapidly to meet changes in demand. The nitrogen left behind after oxygen is consumed
Operational Fuel‐Cell Voltages
at the air electrode can also hinder mass transport at high currents — it effectively
blocks the oxygen supply.
3.8
Combining the Irreversibilities
It is useful to construct an equation that brings together all the irreversibilities associated
with fuel cells. Such an exercise results in the following relationship between operating
voltage and current density:
Vc
Vr
i in r A ln
i in
io
B ln 1
i in
il
(3.22)
where:
●
●
●
●
●
●
●
Vr is the reversible open‐circuit voltage given by equation (2.9), Chapter 2
in is the sum current density equivalent of fuel crossover and the internal current
density, as described in Section 3.5
A is the slope of the Tafel line, as described in Section 3.4.2
io is either the exchange‐current density at the positive electrode if the overpotential
is much greater than that of the negative electrode, or it is a function of both exchange‐
current densities, as given in equation (3.11)
B is the parameter in the mass-transfer overvoltage equation (3.21), as discussed in
Section 3.7
il is the limiting current density at the electrode that has the lowest limiting current
density, as discussed in Section 3.7
r is the ASR, as described in Section 3.6
Example values of the constants are given in Table 3.3 for two different types of fuel
cell.
It is possible to model equation (3.22) by means of a spreadsheet (such as EXCEL), a
program such as MATLAB, or a graphics calculator. It must be borne in mind that there
may be problems at low current densities, as the third term in the equation is only valid
when (i + in) >> io. Also, the equation is not valid when the limiting current density is
exceeded, i.e., (i + in) > il. Given these caveats, the reader should be able to generate graphs
very similar to those displayed in Figures 3.1 and 3.2 by using the data provided in Table 3.3.
Table 3.3 Example values of parameters for equation (3.22).
Parameter
Low temperature
(e.g., PEMFC)
High temperature
(e.g., SOFC)
Vr (V)
1.2
1.0
–2
in (mA cm )
2
2
r (kΩ cm2)
30 × 10−6
300 × 10−6
io (mA cm–2)
0.067
300
A (V)
0.06
0.03
B (V)
0.05
0.08
il (mA cm–2)
900
900
57
58
Fuel Cell Systems Explained
3.9 The Electrical Double-Layer
The inquisitive newcomer to fuel cells is often prompted to ask further about the nature
of the processes occurring at the electrodes. Granted that chemical reactions — oxidation
and reduction — are occurring at the electrodes, it is pertinent to enquire about the nature
of the interaction between the reacting species and the electrode and electrolyte materials
at the molecular or the atomic level. To explore this subject, it is necessary to invoke a
concept known as the ‘electrical double-layer’. First described by Helmholtz as far back as
1853, the concept has helped to explain the properties of many everyday substances from
colloids such as milk or paint to electrical devices such as capacitors and batteries.
Whenever two such different materials are in contact, there is a build‐up of electrical
charge on the surface at the interface between the materials or a charge transfer from
one to the other. In semiconductors, for example, there is a diffusion of positive ‘holes’
and negative electrons across junctions between n‐type and p‐type materials that are in
contact. This forms a ‘double-layer’ at the junction (of electrons in the p‐type region and
‘holes’ in the n‐type) that plays a fundamental role in semiconductor devices such as
diodes, transistors, photosensors and solar cells.
In electrochemical systems, the double-layer forms in part due to diffusion effects
(as in semiconductors) associated with the reactions between the electrons in the
electrodes and the ions in the electrolyte, and also as a result of applied voltages. For
example, the situation depicted in Figure 3.6 might arise at the cathode of a fuel cell with
an acid electrolyte. Electrons will collect at the surface of the electrode, and H+ ions will
be attracted from the bulk to the surface of the electrolyte. The electrons and ions,
together with the oxygen supplied to the positive electrode, will take part in the reaction
given by equation (3.4), i.e.,
Electrode
Electroyte
Figure 3.6 Charge double layer at the cathode surface of a fuel cell.
Operational Fuel‐Cell Voltages
O2 4 e
4H
(3.4)
2H 2 O
The accumulation of positive charge on the surface of the cathode as a result of the
migrated H+ ions and a relatively lower charge in the surrounding electrolyte results in
the formation of an electrical double-layer. The layer has a complex structure with (i) an
inner Helmholtz plane (IHP), which is the layer of absorbed ions on the surface of the
electrode (H+ ions in the case of Figure 3.6), and (ii) an outer Helmholtz plane (OHP),
which represents the position of the ions in the electrolyte closest to the electrode surface. As discussed later in Chapter 4, all the ions in (i) and (ii) are hydrated ions in
PEMFCs. Beyond the OHP, there are ions in the electrolyte that can interact via long‐
range electrostatic forces.
The probability of a reaction taking place depends on the density of the charges,
electrons and H+ ions on the electrode and electrolyte surfaces. Any collection of
charge will generate a difference in electrical potential between the electrode and
electrolyte — this is the ‘activation overpotential’, which was discussed in Section 3.4.
The layer of charge on or near the electrode|electrolyte interface is a store of electrical
energy, and as such behaves much like an electrical capacitor. If the current changes, it
will take some time for the charge (and its associated voltage) to dissipate (if the current
reduces) or build up (if there is a current increase). Consequently, unlike an ohmic loss
in voltage, the activation overpotential does not immediately change with the current.
Consider now the combined effect of the overpotentials on two electrodes of a
complete fuel cell. If the current through the fuel cell suddenly changes, the operating
voltage will show an immediate change due to the internal resistance that is followed by
a fairly slow progress to its final equilibrium value. The behaviour can be modelled by
using an equivalent circuit, with the double-layer represented by an electrical capacitor.
The capacitance of a capacitor, C, is given by the formula:
C
A
d
where ε is the electrical permittivity, A is the surface area
and d is the separation of the plates. For a fuel cell, A is
the real surface area of the electrode, which is several
thousand times greater than its length × width. The
separation, d, is very small, i.e., typically only a few
nanometres. Consequently, the capacitance in some fuel
cells will be of the order of a few Farads, which is high in
terms of capacitance values. (In electrical circuits, a 1‐μF
capacitor is relatively large.) The connection between this
capacitance, the charge stored in it and the resulting
activation overpotential leads to an equivalent circuit, as
shown in Figure 3.7. The resistor Rr simulates the ohmic
losses. A change in current gives an immediate change in
the voltage drop across this resistor. The resistor Ra models
the activation overpotential, and the capacitor ‘smooths’
any voltage drop across this resistor. If the concentration
overpotential were to be included, it would be incorporated
in Ra.
(3.23)
Rr
Ra
E
Figure 3.7 Simple equivalent
circuit model of a fuel cell.
59
60
Fuel Cell Systems Explained
Generally speaking, the capacitance that results from the double layer gives the fuel
cell a ‘good’ dynamic performance, in that the voltage moves gently and smoothly to
a new value in response to a change in current demand. It also permits a simple and
effective way to distinguish between the main types of voltage drop, and hence to
analyse the performance of a fuel cell, as described in the next section.
3.10 Techniques for Distinguishing Irreversibilities
At various points in this chapter, it has been asserted that a certain distinctive type of
overpotential/overvoltage is dominant under a given condition. For example, it has
been said that for an SOFC, the ohmic voltage drop is more important than activation
losses. Much of the evidence that supports this assertion comes from fundamental
experimental measurements. The following describes some of the techniques that are
frequently employed for experimentally characterizing electrochemical cells — first
those relating to individual electrodes and then those that are applied to complete cells.
3.10.1
Cyclic Voltammetry
Cyclic voltammetry (CV) is widely employed in the investigation of electrochemical
reactions on individual electrodes. Most commonly, a three‐electrode cell is used
with a liquid electrolyte, as illustrated in simplified form in Figure 3.8. The setup is
comprised of the following components:
●
●
A ‘working electrode’ that usually consists of a highly-polished, glassy carbon
substrate on which the electrode or catalyst material to be investigated is deposited.
A ‘counter electrode’, usually a flag of platinum of sufficient area to ensure that
any electrochemical reaction occurring at this electrode, does not influence the
performance of the working electrode.
Working electrode
Counter-electrode
Provision for O2 gas
to saturate the
electrolyte solution
Reference electrode
(e.g., Calomel)
Luggin capillary
Cell contains
electrolyte solution
Figure 3.8 Simple 3‐electrode setup for cyclic voltammetry.
Operational Fuel‐Cell Voltages
●
●
A ‘reference electrode’ against which voltage measurements are made; examples are
Pt|H2|H+ (SHE), Hg|Hg2SO4 (mercury/mercurous sulfate), Ag|AgCl|Cl– (silver|silver
chloride) and Hg|Hg2Cl|Cl– (saturated calomel electrode). The reference electrode is
located close to the working electrode, usually via a small Luggin capillary.
Provision may be made for admitting oxygen to the electrolyte solution, e.g., for
performing CV for the ORR on PEMFC catalysts.
The principle of CV operation is as follows. The material of interest, e.g., a carbonsupported platinum catalyst material for the negative electrode of a PEMFC, is prepared
as a fine powder and dispersed in a solvent such as dilute ethanol. A material, e.g., NafionTM,
may be added to facilitate good adhesion to the electrode. The mixture is finely dispersed
by agitation or ultrasonication, deposited on the surface of the working electrode and
then allowed to dry in air. Using, typically, a dilute sulfuric acid solution (0.01–0.1 M) as
the liquid electrolyte, the cell is assembled and the experiment commenced. A potential
difference is applied between the working and reference electrodes and scanned at a
fixed rate towards higher or lower values, as dictated by the reaction of interest. The
current flowing between the working and counter electrodes is recorded as a function
of the applied voltage. Voltage is controlled and measured in CV experiments by means
of a potentiostat, which is an instrument that draws no current from the reference
electrode. When the reaction at the electrode is complete, the voltage is scanned in the
opposite direction (hence the term ‘CV’). If the reaction of interest is reversible, then
the reverse sweep will show this as a current flowing in the opposite direction. The plot
of current versus applied voltage is known as a ‘cyclic voltammogram’; examples will be
examined later in the chapters devoted to low‐temperature fuel cells.
A voltammogram provides information about the oxidation–reduction potential and
the rates of the electrochemical reactions occurring at a given electrode. The technique
is particularly valuable in allowing measurement of the activity of fuel‐cell catalysts
without the need to assemble complete fuel cells or half cells.
The rotating disc electrode (RDE) is an extension of the CV method. This device
employs the same three‐electrode experimental setup as for CV, except that the working
electrode is able to rotate at high speeds. If the electrochemical reaction occurring on
the surface of the electrode is limited by diffusion, this is shown by a change in the
voltammogram as the speed of rotation increases. Above a certain speed, the effects of
diffusion to the surface are minimized. The RDE technique can be employed to probe
the reaction mechanism, for example, to distinguish between 2‐ and 4‐electron transfer
with materials employed as cathode catalysts in PEMFCs. A slight variant of the RDE is
the rotating ring–disc electrode (RRDE), which enables reaction mechanisms to be
elucidated in more detail. Unlike the CV method, in which the electrode is stationary
and the reaction is essentially reversed in the return sweep, this does not take place with
the RDE/RRDE techniques since the surface layer on the catalyst is disturbed by
rotation of the electrode. With rotating electrodes, it is therefore only possible to carry
out linear voltage sweeps and not cyclic sweeps as in CV.
3.10.2
AC Impedance Spectroscopy
AC impedance spectroscopy, sometimes referred to as electrochemical impedance
spectroscopy (EIS), has become a popular means for characterizing both half-cells
and complete fuel cells. In contrast with most other electrochemical methods, this
61
62
Fuel Cell Systems Explained
technique can be applied in situ to a working fuel cell. It is fairly straightforward to
understand in principle, but care must be taken in the analysis of data since many
factors can affect the results. The procedure essentially involves driving a small
variable‐frequency alternating current (AC) through the fuel cell and measuring the
resulting AC voltage across the cell, from which the impedance of the cell can be
determined. Since the AC frequency can be quite low, it is important that the fuel
cell is operating under steady conditions, for example, in the absence of catalyst
activation or deactivation. As with internal resistances, several impedances can
be distinguished in a working cell and attributed to the electrolyte, electrodes and
interfaces.3
The essentials of AC impedance spectroscopy have been known since the 1950s,
but it is really since the emergence of advanced computing systems and frequency
response analyzers (FRAs) in the 1980s that the technique has achieved a routine
status in electrochemistry. The FRA generates a reference voltage sine wave of given
amplitude and frequency, and the magnitude and phase of the resulting AC current
is measured and recorded. Sweeping a range of frequencies gives an ‘impedance
spectrum’, which can be presented as a ‘Bode plot’ of current versus frequency.
The technique is capable of a high degree of precision since unwanted signals in
the spectrum can be filtered out by carrying out measurements over a large number
of cycles.
The measured AC current through the fuel cell is phase‐shifted with respect to the
applied AC voltage sine wave by a phase angle θ. If a radial frequency ω (measured in
radians per second) is defined as:
(3.24)
2 f
where f is the frequency (in Hertz) of the applied voltage, an expression analogous to
Ohm’s law for resistors can be derived, namely:
Z
Et
It
Eo sin t
)
I o sin( t
Zo
sin t
sin t
(3.25)
where Z is the impedance of the system, Et and It are the voltage and current at time
t and Zo is the impedance of the system when both current and voltage are in phase
(θ = 0). It is also possible to show that the impedance can be represented by a complex
number, i.e.,
Z
Zo cos
j sin
(3.26)
Mathematically, this means that Z can be represented by a real and imaginary
component. Plotting the real part (Zre) on the x‐axis and the imaginary part (Zimag) on
the y‐axis of a chart produces a so‐called ‘Nyquist plot’ — an example obtained for a
3 The total internal resistance of a working fuel cell is the sum of the resistances of the various cell
components. It varies with current density, and, as with a battery, the maximum power delivered by a fuel cell
is achieved when the total internal resistance is equal to the sum of the resistances in the external circuit. The
magnitude of the internal resistance is the same as the magnitude of the cell impedance as measured by EIS.
Impedance, however, also has a phase dimension as alternating current is involved.
Operational Fuel‐Cell Voltages
(a)
0.75
0.00
E2
E3
E4
E-2
–Zim /Ω cm2
0.25
E1
0.50
100
0.50 0.75 1.00
1.25 1.50
50
0
0
50
100
150
Zre /Ω cm2
(b)
10–3
10–1
101
103
–75
–50
10
ϕ/degrees
|Z| /Ω cm2
100
–25
1
0
10–3
10–1
101
103
f/Hz
Figure 3.9 (a) Nyquist plot of a practical SOFC anode in 97% CH4 and 3% H2O, at 932°C. Inset: Zoom at
frequencies higher than 1 Hz. (b) Corresponding Bode plots. (Source: Kelaidopoulou, A, Siddle, A, Dicks,
AL, Kaiser, A and Irvine, JTS, 2001, Anodic behaviour of Y0.20Ti0.18Zr0.62O1.90 towards hydrogen
electro‐oxidation in a high temperature solid oxide fuel cell, Fuel Cells, vol 1(3–4), pp. 226–232.)
anode in an SOFC is shown in Figure 3.9a. The advantages of presenting data in this
form are as follows:
●
●
●
Ohmic resistance (Rr) is displayed in the left‐hand region where the semicircle reaches
(or extrapolates to the x‐axis); this represents the point of zero frequency. In the
example under consideration, Rr = 0.48 Ω cm2.
The y‐axis represents the capacitive elements of the cell.
Activation‐controlled processes with distinct time constants show up as unique
impedance arcs, and the shape of the curve provides insight into a possible reaction
mechanism or governing phenomena. In the example shown in Figure 3.9a, two
processes were detected, namely, (i) at high frequency, there was a charge‐transfer
63
Fuel Cell Systems Explained
●
process (Ox + ne− → red), and (ii) at low frequency, careful deconvolution of the curve
distinguished three separate resistive contributions.4
The main disadvantage of the Nyquist plot is that frequency is not directly plotted so
it is difficult to determine the frequency of a point on a Nyquist plot. This can be
overcome by displaying data in the form of a Bode plot, e.g., Figure 3.9b, in which the
impedance (either the real or imaginary component) or phase angle is plotted versus
frequency.
It is possible to fit AC impedance curves to an equivalent electrical circuit made up of
individual resistors and capacitors, such as shown in Figure 3.7. Computer software is
now available that will fulfil this task.
Impedance (Nyquist) plots shown in Figure 3.10a are typical spectra for the ORR at a
cathode platinum|Nafion interface of a PEMFC at different potentials.5 There are two
pronounced arcs, accounting for the charge transfer at high frequencies and masstransfer processes at low frequencies. The charge‐transfer arc decreases at higher
voltages due to the increased rate of the electrochemical reaction, and the arc due
to mass‐transport impedance becomes more dominant. The representative equivalent
(a)
100
–Z″/Ω cm2
64
50
0
0
50
100
150
200
Z′/Ω cm2
(b)
CPE
RΩ
Rct
WS
Figure 3.10 (a) Effect of overpotential on impedance plots for Nafion 117. Applied DC potential: (ο)
0.775 V, (▲) 0.75 V, (∙) 0.725 V and (•) 0.70 V. Temperature 303 K and oxygen pressure 207 kPa. Solid lines
represent fits of the equivalent circuits. (b) Typical equivalent circuit of PEMFC for ORR at Pt/Nafion
interface. (Source: Xie, Z and Holdcroft, S 2004, Polarization‐dependent mass transport parameters
for ORR in perfluorosulfonic acid ionomer membranes: an EIS study using microelectrodes, Journal
of Electroanalytical Chemistry, vol 568, pp. 247–260. Reproduced with the permission of Elsevier.)
4 Kelaidopoulou, K, Siddle, A, Dicks, AL, Kaiser, A and Irvine, JTS, 2001, Methane electro‐oxidation on a
Y0.2Ti0.18 Zr0.62O0.19 anode in a high temperature solid oxide fuel cell, Fuel Cells, vol. 1(3–4), pp. 219–225.
5 Yuan, X, Wang, H, Sun, JC and Zhangm, J, 2007, AC impedance technique in PEM fuel cell diagnosis—A
review, International Journal of Hydrogen Energy, vol. 32, pp. 4365–4380.
Operational Fuel‐Cell Voltages
circuit is depicted in Figure 3.10b to simulate the typical plot, where R, Rct and Ws
represent the ohmic resistance, charge‐transfer resistance and finite‐length Warburg
impedance, respectively. The last-mentioned represents diffusion of the reacting
species (in this case, oxygen). The conventional double‐layer capacitance is replaced by
a constant phase element (CPE) because the capacitance caused by the double‐layer
charging is distributed along the length of the pores in the porous electrode. The circuit
of Figure 3.10b is known as the Randles circuit and is one the simplest commonly used
model of fuel‐cell electrodes.
By carrying out AC impedance on both the anode and the cathode, the user may be
able to determine the contributions made to the overpotentials by individual materials
or processes occurring within the cell. Usually, impedance spectra are more complex
than that illustrated in Figure 3.10; procedures for their interpretation are outside the
scope of this book. Fortunately, AC impedance has become a more popular tool in the
characterization of fuel‐cell systems, and consequently there are many references on
the experimental techniques and interpretation of data.6
3.10.3
Current Interruption
The current‐interrupt technique not only provides accurate quantitative results but
also delivers rapid qualitative indications of internal losses in working fuel cells. Unlike
impedance spectroscopy, it can be performed using standard, low‐cost, electronic
equipment; the basic setup is shown in Figure 3.11. To understand current interruption,
consider a fuel cell that is providing a current at which the concentration (or mass
transport) overpotential is negligible, and, therefore, the voltage drop is caused by
ohmic losses and the activation overpotential. If the current is suddenly cut off, the
charge double layer will take some time to disperse and so will the associated
overpotential. By contrast, the ohmic losses will immediately reduce to zero. The
resulting change in voltage measured at the fuel‐cell terminals when a load is suddenly
disconnected is shown schematically in Figure 3.12.
Fuel cell
A
Digital storage oscilloscope
Figure 3.11 Simple circuit for performing a current‐interrupt test.
6 Example: ‘Basics of Electrochemical Impedance Spectroscopy’, published by Gamry Instruments — available
online at http://www.gamry.com/application‐notes/EIS/basics‐of‐electrochemical‐impedance‐spectroscopy/
65
66
Fuel Cell Systems Explained
Voltage
Slow final rise
to OCV,Va
Immediate rise in
voltage, Vr
OCV– open-circuit voltage
Va – activation overpotential
Vr - voltage rise due to
ohmic losses
Time
Time of current interruption
Figure 3.12 Schematic of voltage against time for a fuel cell after a current‐interrupt test.
Measurement of current interruption is as follows. The switch in the circuit illustrated
in Figure 3.11 is closed, and the load resistor is adjusted until the desired test current is
flowing. The storage oscilloscope is set to a suitable time base, and the load current is
then switched off. The oscilloscope triggering will need to be set so that the instrument
moves into ‘hold’ mode — though with some cells, the system is so slow that the
procedure can be done by hand. The two voltages Vr and Va, shown in Figure 3.12, are
then read off the screen. Although the method is simple, care must be taken when
obtaining quantitative results as it is possible to overestimate Vr by missing the point
where the immediate rise in voltage ends. The setting of the oscilloscope time base will
vary for different types of fuel cell, as determined by the capacitance.
The current‐interrupt test is easy to perform with single cells and small fuel‐cell
stacks. With larger cells and stacks, the switching of high currents can be problematic.
Typical results from three current‐interrupt tests, as shown in Figures 3.13, 3.14 and 3.15,
provide a clear qualitative indication of the importance of the different types of voltage
increases that can be observed. Because oscilloscopes do not show vertical lines, the
appearance of each trace is slightly different from that given in Figure 3.12, namely,
there is no vertical line corresponding to Vr. The tests were performed on three different types of fuel cell: a PEMFC, a DMFC and a SOFC. In each case, the total voltage
drop is about the same (Vr + Va), though the current density certainly is not.
The three examples give a good summary of the causes of voltage losses in fuel cells.
Concentration or mass‐transport losses are important only at higher currents, whereas
in a well‐designed system with a good supply of fuel and oxygen, they should be very
small over the range of cell operating currents. In low‐temperature hydrogen fuel cells,
anode activation can be ignored, and the dominant voltage loss is due to the activation
losses at the cathode, especially at low currents (below about 50 mA cm−2). At higher
currents, namely, above about 50 mA cm−2, the activation and the ohmic losses are similar — see Figure 3.13. In cells using fuels such as methanol, there are considerable
activation losses at both the anode and cathode, and therefore the activation overpotential
Operational Fuel‐Cell Voltages
Figure 3.13 Current‐interrupt test for a
low‐temperature, ambient pressure,
hydrogen fuel cell. Ohmic and activation
overpotentials are similar. (Time scale 0.2 s
per division; i = 100 mA cm−2.)
Va
Vr
Figure 3.14 Current‐interrupt test for a
direct methanol fuel cell. There are large
activation losses at both electrodes so
that, by comparison, the ohmic losses are
barely discernable. (Time scale 2 s per
division; i = 10 mA cm–2.)
Va
Vr
Figure 3.15 Current‐interrupt test for a
small SOFC working at about 700°C. The
large immediate rise in voltage shows that
ohmic losses are responsible for most of the
voltage drop. (Time scale 0.02 s per division;
i = 100 mA cm–2.)
Va
Vr
67
68
Fuel Cell Systems Explained
dominates at all times, as demonstrated in Figure 3.14. On the other hand, the
activation losses become much less dominant in cells that operate at high temperatures,
such as SOFCs at 700°C, and thereby ohmic losses become the main concern, as shown
Figure 3.15.
The aim of the opening three chapters has been to provide a sound understanding of
the general principles of fuel‐cell operation. The following chapters delve more deeply
into the construction, operation and application of the main types of fuel‐cell system.
Further Reading
Büchi, FN, Marek, A and Scherer, GG, 1995, In‐situ membrane resistance measurements in
polymer electrolyte fuel cells by fast auxiliary current pulses, Journal of The
Electrochemical Society, vol 142(6), pp. 1895–1901.
Greef, R, Peat, R, Peter, LM, Pletcher, D and Robinson, J, 2002, Instrumental Methods in
Electrochemistry, Ellis Horwood, Oxford. ISBN‐13: 978‐1898563808.
Hamann, CH, Hamnett, A and Vielstich, W, 2007, Electrochemistry (Second, Completely
Revised and Updated Edition), Wiley‐VCH, Weinheim. ISBN: 978‐3‐527‐31069‐2.
Yuam X‐Zi, Song, C, Wang, H and Zhang, J, 2010, Electrochemical Impedance Spectroscopy
in PEM Fuel Cells, Springer, London. ISBN: 978‐1‐84882‐845‐2 (Print)
978‐1‐84882‐846‐9 (Online).
Zhang, J, (ed.), 2008, PEM Fuel Cell Electrocatalysts and Catalyst Layers, Fundamentals
and Applications, Springer‐Verlag, London. ISBN 978‐1‐84800‐935‐6; DOI
10.1007/978-1-84800-936-3.
Zhang, J, Wu, J, Zhang, H and Zhang, J, 2013 PEM Fuel Cell Testing and Diagnosis,
Elsevier, Burlington, VT. ISBN 978‐0‐44453‐689‐1.
69
4
Proton‐Exchange Membrane Fuel Cells
4.1
Overview
The proton‐exchange membrane fuel cell (PEMFC) — also called the ‘polymer
electrolyte membrane fuel cell’ (PEMFC), ‘solid polymer electrolyte fuel cell (SPEFC)’
and solid polymer fuel cell (SPFC) — is the most widely known version of acid fuel
cell. The design was first developed by General Electric (GE) in the United States for
use by the National Aeronautics and Space Administration (NASA) in the Gemini
manned spacecraft of the 1960s. Instead of the liquid proton‐conducting electrolyte
used in early experimental acid fuel cells, a solid or quasi‐solid ‘membrane’ material
was used. The first PEMFCs employed electrolytes that were based on polymers such
as polyethylene; for instance, the initial NASA fuel cells operated with polystyrene
sulfonic acid. In 1967, DuPont introduced a novel fluorinated polymer based on a
polytetrafluoroethylene (PTFE) structure with the trademark Nafion™ (hereafter,
simply referred to as ‘Nafion’). The PTFE material is used to coat non‐stick cookware
and is highly hydrophobic (i.e., it is not wetted by water). Nafion constituted a major
advance for fuel cells, and it has become an industry standard against which new
polymer membranes are judged. Nevertheless, Nafion is expensive to produce and has
certain limitations, such as the need to be hydrated and therefore is functional only
below about 80°C. For these reasons, many alternative electrolyte materials have been
investigated; the more common varieties are reviewed in Section 4.2.
Apart from the solid oxide fuel cell (SOFC) described later in Chapter 9, the PEMFC
is unique in that it uses a solid sheet of electrolyte that is bound on both sides to sheets
of catalysed porous electrodes. The negative|electrolyte|positive assembly is thus one
item, is very thin and is commonly referred to as the ‘membrane‐electrode assembly’
(MEA). A PEMFC stack comprises several such MEAs connected in series, usually by
means of bipolar plates, as shown in Figure 1.9, Chapter 1. The charge carrier in the
polymer electrolyte is an H+ ion (also known as a proton), and the basic operation of the
cell is essentially as described for the generic acid electrolyte fuel cell illustrated in
Figure 1.3, Chapter 1.
The usual polymer membranes operate at near‐ambient temperatures. This enables
the PEMFC to start up quickly. The absence of corrosive and hazardous fluids that
are present in the electrolytes of alkaline (AFC), phosphoric acid (PAFC) and molten
Fuel Cell Systems Explained, Third Edition. Andrew L. Dicks and David A. J. Rand.
© 2018 John Wiley & Sons Ltd. Published 2018 by John Wiley & Sons Ltd.
70
Fuel Cell Systems Explained
Figure 4.1 Four PEMFC stacks that illustrate development by Ballard Power Systems through the
1990s. The left‐hand stack, the 1989 model, has a power density of 100 W L−1. The right‐hand, 1996
model delivers 1.1 kW L−1 (Source: By kind permission of Ballard Power systems.)
carbonate (MCFC) fuel‐cell systems means that the PEMFC is able to function in any
orientation. Furthermore, the thinness of modern MEAs also enables the production of
compact fuel cells with very high power densities (W L−1). These attributes combine to
make the PEMFC very robust and especially suitable for use in road vehicles and as a
power source for portable electrical and electronic applications.
The early versions of the PEMFC, as used in the Gemini spacecraft, had a lifetime of
only about 500 h, but that was sufficient for those limited early missions. Concern arose,
however, over the reliability of water management in the electrolyte (which is considered
in some detail in Section 4.4), such that NASA selected the AFC for use in the following
Apollo spacecraft. General Electric also chose not to pursue commercial development
of the PEMFC, probably because the costs were seen as higher than those for other types
of fuel cell, such as the PAFC, which was under development for stationary power
applications. At that time, catalyst technology was such that 28 mg of platinum were
required for each cm2 of electrode — compared with 0.2 mg cm−2 or less today.
The development of PEMFCs passed, more or less, into abeyance in the 1970s and
early 1980s. A renaissance of interest began in the latter half of the 1980s and early
1990s, and much of the credit for this must go to Ballard Power Systems of Vancouver,
Canada, and to the Los Alamos National Laboratory in the United States. Developments
in more recent years have brought current densities up to 1 A cm−2 or more, while at the
same time the amount of platinum used in the catalysts has been reduced by two orders
of magnitude. These improvements have led to a huge reduction in cost per kW of
power, and a major increase in power density, as demonstrated in Figure 4.1, which
shows the progress made by Ballard Power Systems during the 1990s.
Both the specific power (W kg−1) and area‐specific power density (W cm−2) of the
PEMFC are higher than for any other type of fuel cell. It is also worth noting that the
2010 performance targets of 650 W kg−1 and 650 W L−1 set by the United States
Proton‐Exchange Membrane Fuel Cells
Figure 4.2 Honda FC stack and Gearbox (exhibited at the Tokyo Motor Show 2007).
Department of Energy (US DOE) for an 80‐kW PEMFC stack were achieved in 2006 by
Honda with a novel 100‐kW vertical‐flow stack that is used in the current FCX Clarity
car. The stack has a volumetric power density of almost 2.0 kW L−1 and a specific power
of 1.6 kW kg−1. In 2008, Nissan also claimed to have achieved 1.9 kW L−1. Since then,
Honda has increased the power density to over 3 kW L−1 (Figure 4.2).1
Proton‐exchange membrane fuel cells are being developed for a very wide range of
operations. For instance, systems as small as a few watts are being marketed for charging mobile phones and other consumer electronic devices, stationary units of several
kW are now in service as for remote telecommunications towers and data centres2,
and others are employed as the power source for domestic‐scale combined heat and
power (CHP) ‘cogeneration’ systems. Nonetheless, their application in road vehicles,
such as cars and buses, has brought the attention of PEMFCs to a greater public.
It could be argued that PEMFCs exceed all other electricity generators in the range of
their possible uses.
In all applications, the three most important distinguishing features of PEMFCs are
as follows:
●
●
●
The type of electrolyte (Section 4.2).
The electrode structure (Section 4.3).
The catalyst (Section 4.3).
1 It is worth cautioning that power density is not the only parameter by which a stack should be judged.
Increasing the current density may improve power density but at the expense of stack lifetime. Therefore, an
important consideration is the required duty and lifetime of the stack for each particular application.
2 Ballard Power Systems has built a 1‐MW stationary power generation system.
71
72
Fuel Cell Systems Explained
Other aspects of system design vary greatly depending on the end use, boundary
conditions and skills of the designer. The most important of these are as follows:
Water management (Section 4.4).
The method of cooling (Section 4.5).
The method of connecting cells in series — bipolar plate designs vary greatly, and
some fuel cells employ markedly different methods (Section 4.6).
The operating pressure (Section 4.7).
The reactants used — pure hydrogen is not the only possible fuel, and oxygen can be
used instead of air (briefly discussed in Section 4.11).
●
●
●
●
●
Some examples of PEMFC systems are examined in Section 4.9.
Over and above technical issues of the PEMFC, cost is the perhaps the most challenging
of the barriers to widespread commercialization. The first commercial PEMFC systems
are now on the market for around US$3000 per kW, i.e., significantly greater than
alternative power sources such as engine‐ or turbine‐based generators. Widely reported
cost targets for stationary PEMFC systems of around US$1000 per kW continue to be a
challenge and are discussed further elsewhere.3
4.2
Polymer Electrolyte: Principles of Operation
4.2.1
Perfluorinated Sulfonic Acid Membrane
Nafion, for many years the industry standard membrane electrolyte in PEMFCs, is a
particular type of perfluorinated sulfonic acid (PFSA). The starting material for Nafion
is the synthetic polymer commonly known as polyethylene or simply polythene. The
molecular structures of ethylene and polyethylene are shown in Figure 4.3.
Polyethylene is modified by chemically substituting fluorine for the hydrogen atoms,
to create a ‘perfluorinated’ polymer. The modified polymer, shown in Figure 4.4, is
H
H
C
C
H
H
H
H
H
H
H
H
H
H
H
H
H
H
H
H
H
C
C
C
C
C
C
C
C
C
C
C
C
C
C
C
H
H
H
H
H
H
H
H
H
H
H
H
H
H
H
Ethylene
Polyethylene (or polythene)
Figure 4.3 Chemical structure of polyethylene.
F
F
C
F
C
F
Tetrafluoroethylene
F
F
F
F
F
F
F
F
F
F
F
F
F
F
F
C
C
C
C
C
C
C
C
C
C
C
C
C
C
C
F
F
F
F
F
F
F
F
F
F
F
F
F
F
F
Polytetrafluoroethylene (PTFE)
Figure 4.4 Chemical structure of PTFE.
3 Staffell, I and Green, R, 2013, The cost of domestic fuel cell micro‐CHP systems, International Journal of
Hydrogen Energy, vol. 38(2), pp. 1088–1022.
Proton‐Exchange Membrane Fuel Cells
PTFE; ‘tetra’ indicates that all four hydrogen atoms in each ethylene group have been
replaced by fluorine. First produced in 1938 and sold by the DuPont Corporation under
the trade name TeflonTM, this remarkable material has exerted a key influence in the
development of fuel cells. The strong bonds between the fluorine and carbon atoms
make PTFE exceptionally resistant to chemical attack and thereby very durable.
Moreover, PTFE is also strongly hydrophobic (i.e., it repels water). Consequently, it is
used in fuel‐cell electrodes to drive product water out of the electrode and thereby
prevent flooding. For the same reason, PTFE is also employed in AFCs and PAFCs.
To make an ion‐conducting electrolyte, PTFE requires further chemical modification,
namely, it has to be ‘sulfonated’. This treatment adds side-chains to the PTFE molecular
backbone, and each of these is terminated with a sulfonic acid (─SO3H) group; there
are several procedures and are mostly proprietary to the membrane manufacturers.
An example of a side-chain structure is given in Figure 4.5 — the details vary both for
different types of Nafion and other PFSAs. In contrast to the creation of side-chains,
the sulfonation of complex molecules is a widely adopted and understood chemical
process. It is used, for example, in the manufacture of detergent. In practice, Nafion is
terminated with side-chains of SO3− ions that are balanced by Na+ ions. In other words,
Nafion may more accurately be considered to be a sodium salt. The ─SO3H groups that
feature for use in PEMFCs are generated by boiling the Nafion with concentrated
sulfuric acid in a final preparative step during which the sodium is discarded as
sodium sulfate.
When the sulfonated polymer is converted to the acidic form, the ─SO3H group is
ionic and so the end of the side-chain is actually an SO3− ion, in which the sulfur atom
is bound to the carbon chain. For this reason, the resulting polymer structure possesses
ionic character and is called an ‘ionomer’. Due to the presence of SO3− and accompanying
H+ ions, there is a strong mutual attraction between the positive and negative ions
from each molecule. Consequently, the side-chains tend to ‘cluster’ within the overall
structure of the material. A key property of sulfonic acid is that it is highly hydrophilic,
that is, it attracts water.4 In Nafion, therefore, the effect is that hydrophilic regions are
Figure 4.5 Chemical structure of a
sulfonated fluoroethylene, also called
‘perfluorosulfonic acid PFTE
copolymer’.
F
F
F
F
F
F
F
F
F
F
F
F
F
F
F
C C C C C C C C C C C C C C C
F
F
F
F
F
F
F
O
F
F
F
F
F
F
F
F C F
F C F
O
F C F
F C F
O S O
O–
H+
4 This is why most detergents are sulfonates. In a detergent molecule, such as an alkylbenzene sulfonate,
the ionic sulfonic end of the molecule mixes readily with water, whereas the polar end of the molecule (the
alkylbenzene) is attracted to the fat, grease and dirt.
73
74
Fuel Cell Systems Explained
Table 4.1 Structure of Nafion and characteristics of other PFSAs.
( CF2CF2)x (CF2CF )y
( OCF2CF )m O
(CF2 )n SO3H
CF3
Structure
parameter
m = 1, x = 5–13.5,
n = 2, y = 1
Trade name and type
Equivalent weight
Thickness (µm)
Nafion 120
1200
260
Nafion 117
1100
175
Nafion 115
1100
125
Nafion 112
1100
80
Flemion ‐ T
1000
120
Flemion ‐ S
1000
80
Dupont
Asashi Glass
m = 0, 1, n = 1–5
Flemion ‐ R
1000
50
m = 0, n = 2–5,
x = 1.5–14
Asashi Chemicals
Aciplex ‐ S
1000–1200
25–100
m = 0, n = 2,
x = 3.6–10
Dow Chemical Dow
800
125
Source: Lee, JS, Quan, ND, Hwang, JM et al., 2006, Polymer electrolyte membranes for fuel cells.
Journal of Industrial Engineering Chemistry, vol. 12(2), pp. 175–183.
created within a generally hydrophobic substance. As mentioned earlier, Nafion is a
specific type of PFSA, and there are many other PFSAs that have been used as fuel‐cell
membranes. Some examples are shown in Table 4.1.
The hydrophilic regions around the clusters of sulfonated side-chains in Nafion and
other PFSAs can lead to the absorption of large quantities of water, which can increase
the dry weight of the material by up to 50%. Within these hydrated regions, the H+ ions
are weakly attracted to the SO3− groups and are therefore mobile; essentially a dilute
acid is created. The resulting material has different microdomains within the macromolecular structure, namely, dilute acid regions, in which H+ ions are attached to water
molecules to create hydronium ions (H3O+), within a tough and strong hydrophobic
structure, as illustrated in Figure 4.6. Although the hydrated regions are somewhat
separate, it is still possible for the H+ ions to move through the supporting long molecule
structure. The proton conductivity of membranes is, nevertheless, higher than would
be expected by simple migration of H3O+ ions. This has led to the view that proton
conduction is via a Grotthus mechanism in which H+ ions move by ‘hopping’ from one
water cluster to the next, a process made easy by the weak hydrogen bonds that have to
Proton‐Exchange Membrane Fuel Cells
Water collects
around the
clusters of
hydrophylic
sulfonate sidechains
Figure 4.6 Structure of PFSA‐type membrane materials: long‐chain molecules containing hydrated
regions around the sulfonated side-chains.
be made and broken with each ion movement. The mechanism was confirmed by
experimental studies reported in 2006.5
For application in fuel cells, Nafion and other PFSA ionomers offer the following
attractive features in that they are:
●
●
●
●
●
Resistant to chemical attack and stable in both oxidizing and reducing environments.
Mechanically strong, on account of the durable PTFE backbone, and so can be made
into very thin films, down to 50 µm.
Acidic.
Able to absorb large quantities of water.
Good proton conductors when well hydrated, to allow H+ ions to move quite freely
within the material.
The ionic conductivity of Nafion depends not only on the degree of hydration, which
is influenced by the temperature and operating pressure, but also on the availability of
the sulfonic acid sites. For example, the conductivity of Nafion membranes quoted in
the literature varies widely depending on the system, pretreatment and equilibrium
parameters used. At 100% relative humidity (RH), the conductivity is generally between
0.01 and 0.1 S cm−1 and falls by several orders of magnitude as the humidity decreases.
Therefore, the degree of hydration has a very marked influence on the ionic
conductivity of the membrane and thereby on performance of the fuel cell. In contrast,
the availability of sulfonic acid sites, usually expressed as the membrane equivalent
weight6 (EW), is relatively unimportant. Values of EW between 800 and 1100 (equivalent
to acid capacities of between 1.25 and ~0.90 mEq g−1) are acceptable for most
5 Tushima S, Teranishi K and Hirai S, 2006, Experimental elucidation of proton-conducting mechanism in
a polymer electrolyte fuel cell by nuclei labelling MRI, ECS Transactions, vol. 3(1), pp. 91–96.
6 Membrane equivalent weight (EW) is defined as the weight of polymer (in terms of molecular mass) per
sulfonic acid group. Ion‐exchange capacity, or acid capacity for PFSAs, is the reciprocal of EW.
75
76
Fuel Cell Systems Explained
membranes because investigations have shown that the maximum ionic conductivity
can be obtained in this range.
It may be also expected that the proton conductivity of a PFSA can be improved by
reducing the thickness of the material. In addition to thickness, however, the proton
conductivity depends on the water content and structural variables such as porosity,
tortuosity, distribution of protons and various diffusion coefficients for the proton
conduction processes. Therefore, whereas making thinner membranes may improve
conductivity, other factors should be taken into consideration such as the fact that
thin materials are inherently less robust and small amounts of fuel crossover can
occur with consequent reduction in the observed cell voltage. For these reasons,
membrane thicknesses of between 80 and 150 µm have been found to be optimum for
most PEMFCs.
Despite being used widely by developers of fuel cells, PFSA membranes suffer from
two major disadvantages, namely: (i) high cost, due to the inherent expense of the
fluorination step in the synthesis of the ionomer and (ii) inability to operate above about
80°C at atmospheric pressure due to evaporation of water from the membrane. With
respect to the latter, higher operating temperatures can be achieved by running the cells
at elevated pressures, but this has a negative effect on system efficiency due mainly to
the additional electrical power required to pressurize the gases. Above 120–130°C, the
PFSA materials undergo a glass transition (i.e., a structural change from an amorphous
plastic phase to a more brittle state) that also severely limits their usefulness.
Membranes that could operate at higher temperatures without the need for
pressurization could therefore bring the following significant benefits:
●
●
●
●
Carbon monoxide concentrations in excess of about 10 ppm at low temperatures
(<80°C) will poison the electrocatalyst used in the Nafion‐based PEMFC. As the
operating temperature increases, so the carbon monoxide tolerance of the platinum
catalyst improves.
Operating at high temperatures has the advantage of creating a greater driving force
for more efficient cooling of the stack. This is particularly important for transport
applications to reduce the need for balance‐of‐plant equipment (e.g., radiators).
High‐grade exhaust heat from high‐temperature fuel cells can be useful for fuel
processing, or in CHP applications.
Increasing the operating temperature will allow the use of a catalyst with lower
activity. Thus the cost penalty associated with expensive platinum catalysts could be
reduced or even avoided.
4.2.2
Modified Perfluorinated Sulfonic Acid Membranes
It was noted earlier that although thin membranes may bring the advantage of reduced
internal ionic resistance, the need to have mechanically strong materials limits how thin
they can be made. To overcome this restriction, some materials — for example, the
Gore Select™ membrane that is favoured by some PEMFC developers — are formed
using a very thin microporous‐based material of expanded PTFE into which an ion‐
exchange resin is incorporated, typically a PFSA or a perfluorinated carboxylic acid.
This technique has enabled membranes with acceptable mechanical properties to be
made as thin as 5–30 µm.
Proton‐Exchange Membrane Fuel Cells
An alternative approach has been to modify chemically the molecular structure of
the polymer so as to increase the porosity at the nanoscale and thereby allow a greater
retention of water. This objective has been achieved by incorporating co‐monomers
with bulky side groups, or by using block copolymers. The resulting membrane material
has high proton conductivity, even at relatively low humidity. Another simple method
for attaining the same outcome is to add a second proton‐conducting material alongside
the polymer. The earliest examples of this approach were the inclusion of small particles
of inorganic, proton‐conducting oxides such as silica (SiO2) or titania (TiO2). Sol–gel
techniques were employed to introduce the oxides with the aim of absorbing water on
the oxide surface so as to limit water loss from the cell via ‘electro‐osmotic drag’.
Unfortunately, such a technique has normally led to significant reduction in the proton
conductivity of the PFSA. Better results have been obtained by incorporating silica‐
supported phosphotungstic acid and silicotungstic acid, zirconium phosphates and silica
alkoxides produced using, e.g., (3‐mercaptopropyl)‐methyldimethoxysilane (MPMDMS).7
4.2.3
Alternative Sulfonated and Non‐Sulfonated Membranes
The high cost of manufacturing the PFSAs has led researchers to seek alternative
membrane materials for PEMFCs (particularly for high‐temperature operation) and also
for application in direct methanol fuel cells (DMFCs). Severe methanol crossover can
occur through traditional PFSA membranes from the anode to the cathode when used in
DMFCs.
Many hydrocarbon polymers have attracted attention in recent years despite the fact
that materials such as phenol sulfonic acid resin and poly(trifluorostyrene sulfonic
acid) were considered during the 1960s but later fell out of favour on the account of
their low thermal and chemical stability. Polymer chemists have evaluated materials
such as trifluorostyrene, copolymer‐based α,β,β‐trifluorostyrene monomer and radiation‐grafted polymer membranes. Of the non‐fluorinated polymers, the most studied
are sulfonated poly(phenyl quinoxalines), poly(2,6‐diphenyl‐4‐phenylene oxide),
poly(aryl ether sulfone), acid‐doped polybenzimidazole (PBI), sulfonated polyether
ether ketone (SPEEK), poly(benzyl sulfonic acid)siloxane (PBSS), poly(1,4‐phenylene),
poly(4phenoxybenzoyl‐1,4‐phenylene) (PPBP) and polyphenylene sulfide.
The range of candidate polymers examined during the 1990s by the Advanced
Materials Division of Ballard Power Systems and by other fuel‐cell companies and
organizations such as the Stanford Research Institute in the United States has been quite
exhaustive. The physical robustness of many polymers can be enhanced via chemical
modification to give increased entanglement of the side chains. Some of these materials
have improved thermal stability, but unfortunately most have generally lower ionic
conductivities than Nafion at comparable equivalent weights. Many others are also
more susceptible than Nafion to oxidative8 or acid‐catalysed degradation.
7 Ladewig, BP, Knott, RB, Hill, AJ, Riches, JD, White, JW, Martin, DJ, Diniz da Costa, JC and Lu, GQ, 2007,
Physical and electrochemical characterization of nanocomposite membranes of Nafion and functionalized
silicon oxide, Chemistry of Materials, vol. 19(9), pp. 2372–2381.
8 Depending on the catalyst used on the oxygen side of the PEMFC, highly oxidizing peroxide species may be
formed. These can attack the membrane and thereby significantly reduce its lifetime. Furthermore, platinum
particles may dissolve in the acid membrane and recombine to form an electrically conductive pathway
through the polymer that serves to reduce the open‐circuit voltage and therefore the cell performance.
77
78
Fuel Cell Systems Explained
L2
N
N
Y
Z
N
N
L1
Figure 4.7 Chemical structure of polybenzimidazole (PBI).
Perhaps the most investigated and representative of the non‐fluorinated hydrocarbon
polymers is PBI (Figure 4.7), which is a heat‐resistant (melting point >600°C), non‐
sulfonated, basic material made by condensation of 3,3′diaminobenzidine and diphenyl
isophthalate. The inherent proton conductivity of PBI is very low, but it is easily doped
with strong acids to form a single‐phase polymer electrolyte with high conductivity.
Phosphoric acid has been found to be most stable and cost‐effective for this purpose. At
high temperatures, it exhibits good thermal stability, adequate mechanical properties
with low gas permeability and low electro‐osmotic drag of water.
A membrane enhanced with phosphoric acid can be prepared either by infusing a cast
film of PBI with phosphoric acid or by actually polymerizing the monomers directly in
polyphosphoric acid (PPA). This acid can then be hydrolysed to phosphoric acid to
provide a membrane of high mechanical stability and with a high loading of phosphoric
acid. Given the latter feature, the proton conductivity approaches that of Nafion
and increases with temperature. When using a phosphoric acid‐doped membrane, a
typical MEA operates in the temperature range of 150–180°C, and the proton conduction
is essentially through the phosphoric acid rather than through water as in the
conventional PEMFC. Because the temperature is so high, water that is produced by
the fuel cell is released as steam, and therefore the pores of the catalyst layer or gas‐
diffusion layer (GDL) are much less likely to be prone to flooding. Phosphoric acid is
allowed to penetrate the catalyst layers (see Section 4.3) and, because it is mobile, care
needs to be taken to ensure that there is sufficient access and egress of the reactant
gases and products on both sides of the fuel cell. The high‐temperature PEMFCs based
around phosphoric acid‐doped polymers require no external humidification and, in
principle, can lead to a much simpler system compared with the traditional design. One
disadvantage, however, is that the catalytic reaction on both electrodes is slower with
phosphoric acid than with PFSAs, and therefore the cell voltage is generally lower for
such high‐temperature PEMFCs. Consequently, catalysts with slightly higher loadings
of platinum are required. Another problem is that although the phosphoric acid is
immobilized on the basic sites of the PBI, the very high loadings cause the polymer to
lose its chemical stability. This shortcoming can be lessened by casting the PBI with a
polyphenolic resin such as polybenzoxazine (PBOA).
High‐temperature polymer membranes for the PEMFC have been developed by
several research groups in numerous universities and companies. Examples are those
investigated by BASF, the Paul Scherrer Institute, and Sartorius (later Elcomax).
Complete high‐temperature PEMFC systems are now also being commercialized by the
Danish company Serenergy for small‐scale stationary power applications (Figure 4.8).
Proton‐Exchange Membrane Fuel Cells
(a)
(b)
Figure 4.8 Serenegy liquid‐cooled high‐temperature PEMFC: (a) stack and (b) system.
(Source: Reproduced with permission of Serenergy.)
4.2.4
Acid–Base Complexes and Ionic Liquids
Two other classes of potentially useful materials for fuel‐cell membranes are acid–base
complexes and ionic liquids. The first category comprises traditional inorganic acids,
such as sulfuric, phosphoric or hydrochloric acid that are embedded within a polymer.
The polymer has to be chemically basic so that the acid is chemically bound within the
structure. Of the many possible complexes, perhaps the one with the most suitable
79
80
Fuel Cell Systems Explained
characteristics for PEMFCs, or especially for DMFCs, is phosphoric acid combined
with PBI or ABPBI (a simpler version of the polymer without the phenylene groups).
The most widely employed ionic liquids in the fuel‐cell industry are the molten alkali
carbonates used in the MCFC, as described in Chapter 8. Although these may be
referred to as ionic liquids, they are usually known as molten salts because at normal
temperatures they are solid and have to be heated above their melting point to have
significant ionic conductivity. The term ‘ionic liquid’ most often is used to describe a
liquid that possesses ionic character but at room temperature. Many organic compounds
fit into this category, and some of these are also being evaluated as a source of possible
membranes for low‐temperature fuel cells. To date, none of these ionic liquids have
progressed from the laboratory into commercial fuel‐cell systems.
4.2.5
High‐Temperature Proton Conductors
As mentioned earlier, phosphoric acid is a good proton conductor and the PAFC is
described in Chapter 5. There are other materials that exhibit proton conduction, but
at much higher temperatures. The most favoured are ceramics with a perovskite
structure,9 notably doped barium and strontium cerates and their mixtures. These
exhibit good proton conductivity (in the order of 10 mS cm−1) in the temperature
range 500–900°C. Unfortunately, due to their basic character, they are unstable in gas
atmospheres that contain H2O, or CO2, H2S, SO2 or SO3, and form Ba(OH)2, BaCO3,
BaS or BaSO4, or the strontium equivalents. The poor chemical stability limits these
materials for application as electrolytes in hydrogen‐only fuel cells.
The barium and strontium cerates are the simplest among many perovskite oxides
that have been investigated for high‐temperature proton conduction, the others being:
●
●
●
II–IV type oxides, e.g., (Ca, Sr, Ba) (Ce, Zr, Ti) O3.
I–V type oxides, e.g., (K Ta O3).
III–III type oxides, e.g., (La Y O3).
The roman numbers refer to groups in the periodic table where the included elements
are to be found. To make such materials proton-conducting, generally they must be
doped with an element of lower valency than the B‐site atom, e.g., Y in BaCeO3, to
increase the concentration of charged species. Some more complex perovskites such as:
●
●
II2–(III/V) type oxides (e.g., Sr2ScNbO6).
II3–(II/V2) type oxides(e.g., Ba3CaNb2O9).
can also be made proton-conducting by making them non‐stoichiometric, e.g.,
Ba3Ca1.18Nb1.82O9‐δ (also known as BCN18).
A range of dopants have been investigated for inclusion in the barium and strontium
cerates (e.g., Y, Tm, Yb, Lu, In or Sc), and it has been found that the larger the ionic radius
or the more basic the dopant, the greater the conductivity for the same dopant level.10
9 Perovskite materials are discussed further in Chapter 9 as many of these materials are also good
oxygen‐ion conductors, suitable as electrolytes for SOFCs.
10 Matsumoto, H, Kawasaki, Y, Ito, N, Enoki, M and Ishihara, T, 2007, Relation between electrical
conductivity and chemical stability of BaCeO3‐based proton conductors with different trivalent dopants,
Electrochem. Solid State Letters, vol. 10, pp. B77–B80.
Proton‐Exchange Membrane Fuel Cells
Alternatives with comparable proton conductivity to the cerium compounds include
fluorite‐related structures, e.g., the tungstates La5.8WO11.7 and La5.7Ca0.3WO11.85, but
their chemical stability has yet to be established. Other materials such as pyrochlores,
e.g., La1.95Ca0.05Zr2O6.975, appear to be more chemically stable, but only show good
proton conductivity up to about 600°C.
Development of high‐temperature proton‐conducting ceramics has been carried out
by groups in the United States under a DOE‐led programme and by a consortium of
research teams in Europe under the EU‐FP7 project. Notably, research in Norway led to
the spin‐out company Protia AS in 2008, with the purpose of commercializing proton
conductors and mixed proton–electron conductors. Applications include hydrogen
separation membranes and enhanced steam reforming systems in addition to proton
ceramic fuel cells. Further to the lanthanum tungstate mentioned earlier, lanthanum
niobate LaNbO4 has also been studied by the researchers in Norway and promises to be
an important material for the future. Another member of the ABO4 family, namely,
lanthanum vanadate LaVO4 was identified by workers in North Europe as a high‐
temperature proton conductor when doped with calcium. Critical to the optimization
of these new materials is the required level of dopants to achieve adequate proton
conduction with good mechanical and chemical stability. To date, conductivity is
significantly lower than alternative oxygen‐ion conductors so that quite thin electrolytes
of the order of 1–10 µm are required.
Although a commercial proton ceramic fuel cell has yet to emerge, by operating up to
900°C platinum‐containing electrode catalysts will clearly not be required. The porous
carbon GDL and carbon‐supported catalyst layer used in PEMFCs will also not feature
in such fuel cells. Rather, the electrodes will need to resemble more closely those in
SOFCs that operate over a similar temperature range. In addition, and in contrast to
SOFCs that feature an oxygen‐ion‐conducting electrolyte, water is produced at the
cathode of the proton ceramic fuel cell. Therefore, although the temperature of
operation may be similar to the intermediate-temperature SOFCs described in
Section 9.1.1, Chapter 9, the system design is likely to be significantly different.
4.3
Electrodes and Electrode Structure
Platinum is the metal with the greatest catalytic activity for both electrode reactions in
the PEMFC. In the early days of the development of this fuel cell, around 28 mg of
platinum was required per cm2 of electrode surface area for each electrode. This high
rate of usage led to the belief, still widely held, that platinum is a major factor in the cost
of PEMFC and that the world’s supply of the metal is not adequate to satisfy the market
for fuel‐cell vehicles should they become widely adopted. Both observations are
misleading. The reality is that platinum usage has been reduced to below 0.2 mg cm−2
and, moreover, yields much better performance in fuel cells today than catalysts of 10
years ago. At such low ‘loadings’, the basic raw material cost at the present prices of
platinum in a 1‐kW PEMFC would be about US$10, so the prospects of mass commercialization appears to have greatly increased. Even so, refinement of catalysts to give
further improvements in performance and lifetime will be necessary for the PEMFC to
achieve widespread commercial acceptance.
81
82
Fuel Cell Systems Explained
Bipolar plate with flow-field channels
for oxidant and fuel
Gas-diffusion layer
Catalyst layers
Electrolyte membrane
Membrane
electrode
assembly
Figure 4.9 Basic structure of a low‐temperature PEMFC with a simple configuration of bipolar plate.
The basic structure of the electrode in different designs of PEMFC is very similar,
despite variations in the details. The negative and positive electrodes are also essentially
the same, and in many PEMFCs they are identical. The main features of a typical planar
PEMFC in which layers of catalyst are sandwiched between the electrolyte membrane
and a porous GDL are shown in Figure 4.9. The GDL, in turn, is in direct contact with
the bipolar flow‐field plate. The following subsections separately provide a description
of each of these components.
4.3.1
Catalyst Layers: Platinum‐Based Catalysts
In a typical PEMFC, the catalyst layer on each GDL has a thickness of around 10 µm and
comprises very small particles of platinum metal on the surface of finely divided carbon
of a somewhat larger particle size. The requirements for the fuel side (negative) and air
side (positive) of the PEMFC are very different. As has been remarked in Section 3.4.2,
Chapter 3, the rate of the oxygen reduction reaction (ORR) is much slower than that of
the hydrogen oxidation reaction. Typically, the exchange‐current density for hydrogen
oxidation is three orders of magnitude higher than that for oxygen reduction, e.g.,
1 mA cm−2 (H2) versus 10−3 mA cm−2 (O2). At a representative operating current density
of 400 mA cm−2, the voltage loss at the anode is about 10 mV, while that at the cathode
is over 400 mV. For this reason, the platinum loading in the catalyst layer on the air
electrode (cathode) is usually much higher than that in the layer on the fuel electrode
(anode).
The carbon in the catalyst layers is usually produced by the pyrolysis of hydrocarbons
to yield a highly porous, nanostructured powder with a high surface area (800–
2000 m2 g−1). An example of such a commercially available powder is Vulcan XC72
(Cabot), which is found in many industrial applications. In the fuel cell, the carbon serves
not only to disperse the active metal but also to provide good electronic conductivity to
enable a high current to be drawn. The method of depositing the platinum on the
carbon usually starts with a precursor solution (e.g., chloroplatinic acid, or another
water‐soluble platinum compound), which is absorbed on the surface of the carbon. The
absorbed precursor may then be chemically reduced (e.g., using sodium borohydride)
or simply heated to decompose the compound and release the metal as finely divided
particles on the surface of carbon clusters. The result of loading the carbon with platinum,
in a somewhat idealized form, is shown in Figure 4.10. This should be compared with
Figure 1.6 in Chapter 1 that shows an electron micrograph of an actual supported
®
Proton‐Exchange Membrane Fuel Cells
Catalyst particles
Carbon support
Figure 4.10 Structure, idealized, of a carbon‐supported platinum catalyst.
catalyst. The platinum is well dispersed on the carbon particles, so that a very high
proportion of the surface area of the metal will be in contact with the gas‐phase
reactants. This high degree of dispersion maximizes the ‘three‐phase boundary’
described in Section 1.3, Chapter 1.
Two alternative methods are generally employed for depositing the catalyst layers in
the MEA. Either the catalyst is first bonded to the appropriate GDL and then to the
electrolyte or it is bonded first to the electrolyte and the GDLs are added afterwards.
The end result is essentially the same in both cases.
Usually, the first requirement is to produce a dispersion of the platinized carbon
powder in a polar and volatile solvent such as ethanol. A small amount of Nafion
solution is normally added to the mixture for reasons that will become apparent later.
PTFE will often be added also to the catalyst layer; during operation of the fuel cell, this
hygroscopic material serves to expel product water to the electrode surface where it
can evaporate. Ultrasonic agitation of the catalyst/ethanol suspension disperses the
powder and creates an ‘ink’ that enables deposition of the catalyst onto the appropriate
cell component (GDL or electrolyte membrane) by a suitable method such as painting,
printing or rolling. The solvent in the ink is allowed to evaporate to leave the solid
catalyst adhering to the given component. If the catalyst is first deposited on the two
GDLs, the resulting two electrodes are then bonded to either side of the polymer
electrolyte membrane by means of the following common procedure:
●
●
●
The electrolyte membrane is cleaned by immersion in boiling 3 vol.% hydrogen
peroxide in water typically for 1 h, and then in boiling sulfuric acid for the same time,
to ensure as full a protonation of the sulfonate groups as possible (and removal of
sodium ions).
The membrane is rinsed in boiling de‐ionized water for another hour to remove any
remaining acid.
The electrodes are placed on the electrolyte membrane, and the assembly is hot
pressed for 3 min at 140°C and high pressure.
The result is a complete MEA.
If the catalyst ink is first deposited directly to the protonated electrolyte rather than
the respective GDLs, then both GDLs must be applied afterwards. This method tends
83
Fuel Cell Systems Explained
(a)
(b)
0.4
0.4
0.2
0.2
0.0
ji/mA cm–2
ji/mA cm–2
84
1
–0.2
2
–0.4
3
4
6
0.0
1
–0.2
2
3
4
6
–0.4
–0.6
–0.6
5
5
–0.8
–0.8
–0.2 0.0
0.2 0.4 0.6
E/ V vs. SCE
0.8
1.0
1.2
–0.2
0.0
0.2 0.4 0.6
E/ V vs. SCE
0.8
1.0
1.2
Figure 4.11 Cyclic voltammograms for thin‐film platinum electrodes (curves 1–5) and bulk platinum
(curve 6) in (a) argon‐saturated 0.1 M HClO4 and (b) 0.05 M H2SO4. Film thickness: (1) 0.25, (2) 0.5, (3) 1,
(4) 2 and (5) 10 nm. Sweep rate: 100 mV s−1. Note the negative peak at 0.40–0.45 V is caused by the
ORR. (Source: Adapted from Thompsett, D, 2003, Catalyst for the proton‐exchange membrane fuel
cell, Chapter 6, in Fuel Cell Technology Handbook, CRC Press, Boca Raton, FL, ISBN: 978‐1‐4200‐4155‐2.)
to result in a thinner catalyst layer and may be preferred for some applications, but
otherwise the MEA gives similar results to that produced by the alternative procedure
outlined earlier.
Both of the procedures for assembling a PEMFC, while being low cost and amenable
to volume production, have the disadvantage of producing relatively thick layers of
catalyst in which platinum is underutilized. More recently, other means of depositing
the active metal onto carbon have been investigated with a view to improving its
effectiveness. Emerging methods include various modified thin‐film techniques,
electrodeposition and sputter deposition, dual ion beam‐assisted deposition, electroless
deposition, electrospray processes, and direct deposition of platinum sols. For example,
platinum particles of less than 5 nm in diameter can be plasma sputtered directly onto
carbon nanofibres11 to produce a catalyst with a loading of between 0.01 and 0.1 mg cm−2.
The performance of platinum catalysts depends very much on the active surface area,
i.e., on the degree of dispersion and the particle sizes. Cyclic voltammograms for
thin films of platinum on carbon are displayed in Figure 4.11. The data show that the
intensity of the oxygen reduction peak increases for film thicknesses between 2 and
10 nm. This complies with data published elsewhere that suggest the optimum size of
platinum particles supported on carbon for catalysing the ORR is between 2 and 4 nm.
Although carbon blacks have been well proven and continue to be used in practical
PEMFCs, both single‐walled and multiwalled carbon nanotubes, as well as graphene,
have been investigated recently as alternatives. The single disadvantage of these forms
of carbon is that they all have intrinsically low surface areas, which is a feature that does
not favour the production of a very active catalyst. On the other hand, the highly ordered
surfaces of carbon nanotubes and graphene do appear beneficial to some non‐precious
metal catalysts, as discussed in the next section.
11 Caillard, A, Charles, C, Boswell, R and Brault, P, 2008, Improvement of the sputtered platinum
utilization in proton exchange membrane fuel cells using plasma‐based carbon nanofibres, Journal of
Physics D‐Applied Physics, vol. 41(18), pp. 1–10.
Proton‐Exchange Membrane Fuel Cells
Cyclic voltammetry and especially the characterization of catalysts using rotating disc
electrodes have identified that two fundamental reactions can occur at the positive
electrode in the PEMFC. The first is the more normal oxygen reduction via a 4‐electron
transfer process, namely:
O2 4 H
4e
2H 2 O
(4.1)
The second reaction is via a 2‐electron transfer intermediate reaction as follows:
O 2 2H
2e
H2O2
(4.2)
The peroxide reaction (4.2) is favoured at cathode potentials of less than 0.5 V with
respect to the standard hydrogen electrode. Peroxide formation may also occur if
hydrogen can crossover (see Section 3.3, Chapter 3) through the membrane and then
become oxidized directly on the cathode. Peroxide reacts with the electrolyte and can
accelerate electrode degradation. It is important therefore to ensure that in the
PEMFC crossover is minimized and that the electrode potentials are maintained
within safe limits.
4.3.2
Catalyst Layers: Alternative Catalysts for Oxygen Reduction
The high cost of platinum has spurred researchers both to reduce its loading in
catalysts and to seek cheaper alternatives. The reason platinum is so active for both
hydrogen oxidation and oxygen reduction has puzzled chemists for many years.12
Current understanding is that the high activity arises partly because the metal loosely
adsorbs molecules such as oxygen or hydrogen on its surface and also because it can
ease the dissociation or splitting of the adsorbed molecules into adsorbed atoms, which
are then able to react.13 The strength of the chemical bond between oxygen atoms and
atoms on the metal surface is dependent on the crystal plane and edges exposed. The
oxygen–metal bond strength can also be influenced by alloying platinum with other
metals. To this end, nickel, rhodium, iridium, cobalt and other transition elements have
been combined with platinum to promote the dissociative adsorption of oxygen. The
influence of combining different metals with platinum has been calculated from first
principles, and experiments have verified that metals such as ruthenium have positive
effects on the surface properties of platinum. A more recent development has involved
depositing the platinum as a shell of single‐atom thickness (a monolayer), or even as
small islands a few atoms thick, on particles of other metals such as ruthenium or
rhodium. These procedures increase the dispersion of the active metal and potentially
reduce the cost of the catalyst.
To lower the costs of the catalyst further, other materials have been evaluated that do
not involve platinum or platinum‐group metals. Over the past 10 years, the alternatives
that have received the most attention are as follows.
12 Platinum is very active for hydrogen oxidation but less active for oxygen reduction in the PEMFC. As
will be shown, other metals work well for oxygen reduction in the alkaline fuel cell, where the reduction
reaction mechanism is somewhat different.
13 Holton, OT and Stevenson, JW, 2013, The role of platinum in proton exchange membrane fuel cells,
Platinum Metals Rev., vol. 57(4), pp. 259–271.
85
86
Fuel Cell Systems Explained
4.3.2.1
Macrocyclics
Transition metal macrocyclic compounds have been examined as potential ORR
catalysts since the early 1960s. The compounds have a molecular structure in which a
central transition metal atom is enclosed within a much larger cyclic organic molecule.
Often the metal atom is linked to nitrogen atoms, and a common characteristic is the
MN4 structure, in which the metal atom is bound to four nitrogen atoms. The structure
is an example of chelation, and therefore the molecule is also known as a ‘chelate’.
Among the many series of macrocyclic compounds, phthalocyanines (Pc) complexed
with various transition metals such as iron, cobalt, nickel and copper have been
thoroughly investigated as oxygen reduction catalysts. Phthalocyanines have been
known since the beginning of the 20th century and are widely used as dyes. Of the
various phthalocyanines evaluated for the reduction of oxygen in fuel cells, the
complexes with cobalt and copper appear to be the most stable, whereas those with iron
and cobalt seem to have the best combination of activity and stability.
The chelates of 5,14‐dihydro‐5,9,14,18‐dibenzotetraaza[14]annulene, or ‘tetraazaanulene’ (TAA) (see Figure 4.12) are another class of macrocyclic complex that exhibits
good potential for ORR catalysis. Porphyrins are the second major group of macrocyclics that have been considered as non‐precious metal catalysts; some examples are
tetraphenylporphyrin (TPP) and tetramethoxyphenyl‐porphyrin (TMPP), also shown
in Figure 4.12. In many cases, these are absorbed onto the carbon carrier, which is then
heated to a high temperature (typically, 800–900°C) to decompose the porphyrin
molecule and thereby allow the metal to bind directly to nitrogen that has, in turn,
become bound to the carbon surface. Indeed, functionalization of the carbon support
surface with nitrogen only (e.g., by treating with concentrated nitric acid) has been
found to enhance the activity of TPP and TMPP catalysts. The study of macrocyclic
compounds as ORR catalysts is currently a very active field of research.
N
N
M
N
N
OMe
TAA
NH
N
NH
N
MeO
N
HN
TPP
OMe
N
HN
OMe
TMPP
Figure 4.12 Molecular structures of macrocyclic organic frameworks used for ORR catalysts:
tetraazaanulene (TAA), tetraphenylporphyrin (TPP), tetramethoxyphenyl‐porphyrin (TMPP).
Proton‐Exchange Membrane Fuel Cells
4.3.2.2
Chalcogenides
In the context of ORR catalysts, ‘chalcogenides’ refer to either sulfides or selenides of
various transition metals. Among the many examples examined to date, Co3S4 and
CoSe2 supported on carbon, as well as various ternary variants such as W–Co–Se, are
all claimed to have high ORR activity.
4.3.2.3
Conductive Polymers
Polymers such as polyaniline (pani), polypyrrole (Ppy) and poly(3‐methylthiophene)
(P3MT) can be used to prepare electronically conducting materials in which metal
atoms, e.g., iron, cobalt or nickel are bonded to the nitrogen atoms within the polymer.
Several of these have been shown to have appreciable catalytic activity for oxygen
reduction.
4.3.2.4
Nitrides
Building on the notion that nitrogen is required for macrocyclics and conductive
polymers to be endowed with catalytic activity, some researchers have explored the
prospects of transition metal nitrides. Whereas tungsten and molybdenum nitrides
supported on carbon exhibit some promise, it has yet to be determined whether these
candidates can be engineered with sufficient activity and longevity to compete with the
established platinum catalysts.
4.3.2.5
Functionalized Carbons
There are essentially two ways by which carbons can be treated or ‘functionalized’14 to
achieve nitrogen on the surface. The first is simply by treatment with nitric acid or by
heating the carbon in nitrogen or ammonia. The process can be conducted either before
the active metal is added or after it is impregnated by a salt of the appropriate metal, e.g.,
cobalt acetate or nickel nitrate (acetate and nitrates are easily decomposed by heating, to
leave the metal atoms bound to the carbon). The second and widely adopted method of
functionalizing carbon is to use a transition metal complex as the source of metal that
when decomposed will ensure a high dispersion of metal atoms on the carbon surface.
Perhaps the most frequently chosen complex has been 2, 4, 6‐tris(2‐pyridyl)‐1,3,5‐triazine
(TPTZ). Typically, the metal‐TPTZ complex is impregnated
into the porous carbon, carbon nanotube or graphene and then
decomposed by heating in the absence of air. Highly active
N
catalysts have been obtained by this procedure. It has been found
that the activity is dependent on the experimental conditions
N
N
(e.g., temperature, heating rate), as well as on the type of
N
N
carbon. Interestingly, the more ordered the carbon structure,
N
the higher the activity of the catalyst is. Catalysts prepared using
Fe‐TPTZ (Figure 4.13) for example, can approach the same
activity as platinum‐based catalysts, but these are supported
Figure 4.13 Molecular
on expensive graphene materials and no doubt much work will structure of 2,4,6‐tris
have to be undertaken if they are to compete with platinum‐ (2‐pyridyl)‐1,3,5‐triazine
(TPTZ).
based materials in terms of long‐term activity in fuel cells.
14 Functionalization, broadly, is the addition of functional groups onto the surface of a material by chemical
synthesis methods. A functional group is a small number of atoms or bonds within a molecule that
determines the chemical properties of the group and of the molecule to which it is attached.
87
88
Fuel Cell Systems Explained
4.3.2.6
Heteropolyacids
Indicating that the field of novel ORR catalysts is by no means exhausted, mention
should be made of recent work on a particular class of inorganic compounds known as
heteropolyacids. Some of these, such as H3PMo12O40 and H3PW12O40, have received
particular attention due to their acidic and redox properties, stability at elevated
temperatures, commercial availability and relative ease of synthesis. Heteropolyacids are
also proton conductors — a feature which may be exploited in the design of the PEMFC.
4.3.3
Catalyst Layer: Negative Electrode
As mentioned earlier, there is less incentive for developers to seek non‐platinum‐based
catalysts for the negative electrode of the fuel cell because less platinum is required for the
hydrogen oxidation reaction. On the other hand, the anode catalyst is susceptible to
poisoning by sulfur and CO, both of which may be present in the hydrogen fed to the fuel
cell, particularly if it has been produced from hydrocarbons. If CO is in the fuel entering
the fuel cell at a concentration of more than a few ppm, it will preferentially adsorb on
the surface platinum atoms and reduce the activity of the catalyst. If the partial pressure of
CO in the fuel stream is low (i.e., below a few ppm), its adsorption on the anode catalyst is
reversible. In such a situation, the catalyst can be kept active, for example, by regularly
purging the fuel side with a small amount of oxygen or briefly applying a negative potential
to the electrode. This technique has been applied in some practical fuel‐cell systems.
Another method of increasing the allowable concentration of CO on the negative electrode
is to use an alloy of platinum and ruthenium, rather than simply platinum, as the catalyst.
4.3.4
Catalyst Durability
Early PEMFCs had lifetimes that were limited not only by the stability of the membrane
but also by the durability of the catalysts. Over the past 10 years, remarkable progress
has been made in the understanding of catalyst durability and has resulted in a significant increase in the expected lifetime. Catalyst degradation is now known to occur,
variously, through the sintering of platinum particles, dissolution of platinum and
corrosion of the carbon support. Sintering of platinum particles on the carbon support
decreases the catalytically-active surface area. It may take place by a dissolution–
precipitation mechanism in which small metal particles of the catalyst may dissolve into
the acidic operating environment and then precipitate onto larger metal particles and
thus promote particle growth, or the particles may directly coalesce with each other due
to movement on the carbon surface. Both mechanisms occur to some degree, with
dissolution–precipitation being more prevalent when load changing or shutdown/
start‐up occurs. The sintering of the catalyst may be reduced by strengthening the
interaction between the catalyst metal and the supporting carbon. For example, grafting
polyaniline to the carbon surface has been found to decrease the mobility of the metal
particles. The nitrogen moiety in the polyaniline has a lone pair of electrons that can
anchor the platinum particles. Dissolution of platinum is accelerated by voltage cycling,
which, for example, is experienced in a fuel‐cell vehicle where acceleration and braking
give rise to a varying load on the PEMFC stack. One way of reducing the effect is to
hybridize the fuel‐cell system in a vehicle with a battery.
Corrosion of the carbon support is generally not an issue for fuel‐cell systems that
operate at steady-state but becomes an important degradation mechanism with repeated
Proton‐Exchange Membrane Fuel Cells
start–stop cycles. When a fuel supply is shut off, air may leak into the anode compartments of a stack and cause a hydrogen–air front to form in the flow-fields of the
separator plates. The result is spatially separated oxygen and hydrogen on the negative
side of the cells that can cause a lowering of the potential at the positive electrodes,
leading to oxidation of carbon in the cathode GDL and/or catalyst layers. Corrosion of
carbon on the cathode side results in a thinning of the catalyst layer that can be
mitigated by changing the type of carbon in the catalyst to a more stable graphitic form
than the conventional pyrolytic variety, by purging hydrogen quickly from the negative
side of the cell on shutdown or by purging oxygen from the positive side.15
4.3.5
Gas‐Diffusion Layer
Commercial GDLs are made of porous conductive material — usually carbon fibres — in
the form of paper or thin fabric/cloth and typically have a thickness of 100–400 µm.
‘Gas‐diffusion layer’ is a slightly misleading name for this part of the electrode, as it does
much more than simply provide a porous structure so that reactant and product gases
can diffuse respectively to and from the catalysts. Namely, it also forms an electrical
connection between the carbon‐supported catalyst and the bipolar plate, or other
current‐collector. In addition, the GDL carries the product water away from the electrolyte
surface and forms a protective layer over the very thin layer of catalyst.
The structure of the electrolyte | catalyst layer and the GDL is shown in idealized
form, in Figure 4.14. The carbon‐supported catalyst particles are joined to the electrolyte
Platinum particles
supported on carbon
Gas-diffusion layer,
e.g., carbon cloth fibres
Main bulk of electrolyte
Figure 4.14 Simplified structure of a PEMFC electrode.
15 Hydrogenics has a patented method of removing oxygen from the positive electrode on shutdown that
involves sealing the supply of air to the cathode of the fuel‐cell stack and electrochemically reducing the
remaining oxygen by using a buffer of hydrogen adjacent to the anode that is reserved for this purpose.
89
90
Fuel Cell Systems Explained
A thin layer of electrolyte
polymer adheres to catalyst
metal particles, promoting
the three-phase contact
between electrolyte,
reactant gas (hydrogen at
the anode and oxygen at the
cathode) and the catalyst
surface
Main bulk of the
electrolyte
Figure 4.15 Enlargement of part of Figure 4.14 to show that the electrolyte reaches out to
the catalyst particles.
on one side, and the GDL (current‐collecting, water‐removing, physical support) on the
other. The hydrophobic PTFE that is needed to remove water from the catalyst is not
shown explicitly in the figure, but will almost always be present.
Two further points need to be made. The first relates to the impregnation of the
electrodes with electrolyte material. A section of the catalyst–electrode region is shown
enlarged in schematic form in Figure 4.15. The electrolyte material extends out to the
catalyst particles sufficiently to provide proton transport to and from the catalyst, which
is where the electrode reactions take place. An important point is that only the catalyst
which is in direct contact with both the electrolyte and the reactant gas can be active for
the electrochemical reactions at the electrodes (reactions only occur at the three‐phase
boundary). To maximize catalyst activity, therefore, the catalyst layer of each electrode
is lightly covered with electrolyte, usually by brushing the surface with a solubilized form
of the electrolyte. In the case of the ‘separate electrode’ method of MEA preparation, this
procedure is conducted before the electrode is hot pressed onto the membrane. By
contrast, the alternative ‘integral membrane–electrode’ process is undertaken before
the GDL is added.
The second point relates to the selection of the GDL, which is generally either a carbon paper or carbon cloth material. Carbon paper (e.g., Toray paper) is chosen when it
is required to make the cell as thin as possible in compact designs. Such paper, which is
made by pyrolysing a non‐woven carbon fibre sheet, has good conductive properties
but tends to be brittle and fragile. By virtue of their greater thickness, carbon cloths will
absorb a little more water than paper and thereby will be less prone to flooding. Cloths
also simplify the mechanical assembly, since they are inherently more flexible and
will deform under compressive forces. Consequently, cloths can fill small gaps and
irregularities in the manufacture and assembly of bipolar plates. On the other hand,
cloths may slightly deform out into the gas‐diffusion channels on the bipolar plates and
thereby may restrict gas flow through the channels. The Elat range produced by Nuvant
Systems Inc. is a commercial cloth that is a popular choice for GDLs. The cloth is
prepared by loading highly conductive carbon fibres with carbon black.
Another GDL innovation is the addition of a very thin microporous hydrophobic
layer of carbon between the GDL and the catalyst layer. The layer consists principally of
carbon black with a 10–40 wt.% loading of PTFE as determined by the type of carbon.
Carbon plays an important role in both the catalyst layer and the GDL of both the
PEMFC and, as shown later, the PAFC. It is not surprising, therefore, that the discovery of
®
®
Proton‐Exchange Membrane Fuel Cells
Figure 4.16 Example of a membrane electrode assembly (MEA). The membrane is a little larger than
the electrodes that are attached. The 10 cm2 membrane is typically 0.05–0.1 mm thick, the electrodes
are about 0.03 mm thick, and the GDL is between 0.2 and 0.5 mm thick.
carbon nanotubes and graphene at the end of the 20th century stimulated considerable
research activity for their application in fuel‐cell systems.16 In the case of the carbon
GDL, the type of carbon, its structure, thickness and electrical conductivity are all
influenced by cell operating conditions. For example, the GDLs in conventional low‐
temperature PEMFCs must be able to accommodate water movement as discussed in
Section 4.4. They tend therefore to have an open‐pore structure to allow for water
diffusion out of the cell. By contrast, the GDLs in PEMFCS operating at high temperatures
(above 100°C) do not need to have such open porosity as liquid water is less likely to
condense out and flood the electrodes.
In summary, the MEA is the key component of a PEMFC and irrespective of its
method of manufacture, every MEA possesses positive and negative electrodes, each of
which incorporates catalyst material. A practical 10‐cm2 MEA is shown in Figure 4.16.
The nature of the electrolyte may differ according to the operating temperature of the
stack. Also important is the way in which the MEA is incorporated in the construction
of fuel‐cell stacks. The design does vary significantly between manufacturers and is
influenced by the application as discussed in the sections that follow.
16 Dicks, AL, 2006, The role of carbon in fuel cells, Journal of Power Sources, vol. 156(2), pp. 128–141.
91
92
Fuel Cell Systems Explained
4.4 Water Management
4.4.1
Hydration and Water Movement
It will be clear from the description of proton‐exchange membranes given in Section 4.2
that, particularly for the PFSA versions, there must be sufficient water in the polymer
electrolyte to maintain high proton conductivity. At the same time, the water content
must be managed to prevent flooding in either the catalyst layer or the GDL.
In an ideal PEMFC, the water that forms at the positive electrode would be expected
to keep the electrolyte at the correct level of hydration. Air would be blown over the
electrode and, as well as supplying the necessary oxygen, it would dry out any excess
water. Because the membrane electrolyte is so thin, water would diffuse from the
positive side to the negative, and throughout the whole electrolyte a suitable state of
hydration would be achieved without any special difficulty. This preferred situation can
sometimes be achieved but relies on a good engineering design.
There are several complications. During operation of the cell, the H+ ions moving
from the negative to the positive electrode pull water molecules with them — a
process usually referred to as ‘electro‐osmotic drag’. Typically, between 1 and 2.5 water
molecules are conveyed for each proton. This means that, especially at high current
densities, the negative side of the electrolyte can become dried out, even if the positive
side is well hydrated. Another major problem is the drying effect of air at high
temperatures; this issue is discussed quantitatively in Section 4.4.2. Suffice it to say, at
temperatures of over about 60°C, the air will always dry out the electrodes faster than
water is produced by the hydrogen–oxygen reaction. It is customary to keep the
membrane sufficiently hydrated by humidifying the air, or the hydrogen, or both,
before entry into the fuel cell. Such action may seem bizarre, as it effectively adds
by‐product to the reactants, but sometimes it is necessary and, moreover, it can greatly
improve the performance of the fuel cell.
To achieve uniform proton conductivity throughout the whole fuel cell, the degree of
hydration in the electrolyte must similarly be uniform. In practice, some parts may be
correctly hydrated, others too dry and others overhydrated or flooded. For example,
consider that air entering the cell may be quite dry, but by the time it has passed over
some of the electrode, it may have achieved the optimum level of humidity. On reaching
the exit, however, the air may be so saturated that it is unable to remove any more of the
water produced. This is a particular problem when designing larger cells and stacks.
The different water movements are shown in Figure 4.17; fortunately they are predictable and controllable.17 For example, water productions at the cathode and electro‐
osmotic drag are both directly proportional to the current. Both water generation and
drag can lead to build-up of liquid water at the cathode of an operating cell. This build‐up
creates a driving force for back diffusion of water from the cathode to the anode,
which helps to keep the membrane uniformly hydrated. The water lost from the cell by
evaporation is governed by the RH of the gases on either side of the cell and can also be
17 Further quantitative discussion of the various water fluxes through the polymer membrane is provided
in: Kumbur, EC and Mench, M, 2009, Water management, in Garche, J, Dyer, C, Moseley, P, Ogum, Z, Rand,
DAJ and Scrosati, B (eds.), Encyclopedia of Electrochemical Power Sources, vol. 2, pp. 828–847, Elsevier,
Amsterdam.
Proton‐Exchange Membrane Fuel Cells
Anode
Electrolyte Cathode
H2O
Water may
back diffuse
from the
cathode to the
anode, if the
cathode side
holds more
water
Water maybe
supplied by
externally
humidifying the
hydrogen
supply
Water will be
produced
within the
cathode
Water will be
dragged from the
anode to the
cathode side by
protons moving
through the
electrolyte
Water will be
removed by the
O2-depleted air
leaving the fuel
cell
Water may be
supplied by
externally
humidifying
the air/O2
supply
Water may be
removed by
circulating
hydrogen
Figure 4.17 Schematic illustration of the different water movements to, within and from the
electrolyte of a PEMFC.
predicted with care, using the theory outlined in the following Section 4.4.2. Therefore
external humidification of the reactant gases prior to entry to the fuel cell, if employed, can
be controlled and can also help to achieve uniform humidification throughout the MEA.
As operating experience with PEMFC systems has grown, so has the understanding
that excessive water in the MEA not only has an immediate negative effect on the cell
performance but also can have long‐term damaging effects. These are caused by the
following actions:
●
●
Internal stresses within the electrolyte. Since water causes swelling of the membrane,
uneven distribution throughout the electrolyte can invoke physical stresses and
degradation of the electrolyte and catalyst layers.
Contamination through water‐soluble ionic species. Post‐service analysis of fuel cells
has shown appreciable accumulation of calcium, iron oxides, copper magnesium and
other metals.
93
94
Fuel Cell Systems Explained
●
Freeze‐out damage. If a small quantity of liquid water remains present in the GDL or
catalyst layers when a PEMFC stack is shut down and the temperature falls below the
freezing point, permanent damage can be caused to these layers. It is essential to
purge any excess water from the PEMFC to prevent such an occurrence.
4.4.2
Air Flow and Water Evaporation
Except for the special case of PEMFCs supplied with pure oxygen, it is universally the
practice to remove the product water via the air that flows through the cell. Consequently,
the air will always be fed through the cell at a rate faster than that needed just to supply
the necessary oxygen. If it were fed at exactly at the ‘stoichiometric’ rate, there would be
substantial ‘concentration losses’, as described in the Section 3.7, Chapter 3, because the
exit air would be completely depleted of oxygen. In practice, the stoichiometry (λ) will
be at least 2. Section A2.2, Appendix 2, provides the derivation of the useful equation
(A2.10), which relates the air flow rate, the power of a fuel cell and the stoichiometry.
Problems arise because the drying effect of air has a non‐linear relationship with
temperature. To understand this characteristic, consideration must be given to the
precise meaning and quantitative effects of terms such as ‘RH’, ‘water content’ and
‘saturated vapour pressure’.
The partial pressures of the various gases that make up air have been given in
Section 2.5, Chapter 2. The analysis, however, ignored the fact that air also contains
water vapour. A straightforward way of measuring and describing the amount of water
vapour is to give the ratio of water to the other gases, namely, nitrogen, oxygen, carbon
dioxide and others that make up ‘dry air’. This quantity is usually given the symbol ω and is
known variously as the ‘humidity ratio’, the ‘absolute humidity’ or the ‘specific humidity’;
it is defined as:
mw
ma
(4.3)
where mw is the mass of water present in the mixture and ma is the mass of dry air, i.e.,
the total mass of the air is mw + ma.
The humidity ratio does not, however, give a very good idea of the drying effect, or the
‘feel’, of the air. Warm air with quite high water content can feel very dry and indeed
have a very strong drying effect. On the other hand, cold air with low water content can
feel very damp. This characteristic is due to changes in the ‘saturated vapour pressure’
of the water vapour, which is the partial pressure of the water when a mixture of air and
liquid water is at equilibrium, i.e., when the rate of evaporation of water in the air is
equal to the rate of condensation. When the air cannot hold any more water vapour at
a given temperature and pressure, it is said to be ‘saturated’.
Air that has no ‘drying effect’ is fully saturated with water and could reasonably be
said to be ‘fully humidified’. This state is achieved when P w = Psat, where P w is the partial
pressure of the water and Psat is the saturated vapour pressure of the water. The ratio of
these two pressures is the ‘RH’, namely:
Pw
Psat
(4.4)
Proton‐Exchange Membrane Fuel Cells
Typical relative humidities vary from about 0.3 (or 30% RH) in the ultra‐dry conditions
of the Sahara desert to about 0.7 (or 70% RH) in a city such as Brisbane or New York on
an ‘average day’. Very important for fuel cells is the fact that the drying effect of air, or
the rate of evaporation of water, is directly proportional to the difference between the
water partial pressure P w and the saturated vapour pressure Psat.
The complication for PEMFCs is that the saturated vapour pressure varies with
temperature in a highly non‐linear way, i.e., Psat increases more rapidly at higher
temperatures. The saturated vapour pressure for a range of temperatures is listed in
Table 4.2. Given the rapid rise in P sat with temperature, air that might be only
moderately drying (say 70% RH) at ambient temperature can be fiercely drying when
heated to about 60°C. For example, for air at 20°C and 70% RH, the pressure of the
water vapour in the mixture is:
Pw
0.70 Psat
0.70 2.338 1.64 kPa
(4.5)
If this air is then heated to 60°C, at constant pressure, without adding water, then Pw
will not change and so the new RH will be:
Pw
Psat
1.64
19.94
0.08 8%
(4.6)
This is very dry — more extreme than in the Sahara desert, for example, where the RH
is typically about 30%. Such a condition would have a catastrophic effect on polymer
electrolyte membranes, which not only require high water content but also are very thin
and therefore prone to drying out rapidly.
The ‘dew point’ is an alternative way of describing the water content. This is the
temperature to which the air should be cooled to reach saturation. For example, if the
partial pressure of the water in air is 12.35 kPa, then, referring to Table 4.2, the dew
point would be 50°C.
As indicated in the previous section, it is sometimes necessary to humidify the gases
going into a fuel cell to ensure an adequate level of hydration throughout the electrolyte
Table 4.2 Saturated vapour pressure of water at selected
temperatures.
T (°C)
Saturated vapour
pressure (kPa)
15
1.705
20
2.338
30
4.246
40
7.383
50
12.35
60
19.94
70
31.19
80
47.39
90
70.13
95
96
Fuel Cell Systems Explained
membrane. To do this in a controlled way might involve calculation of the mass of water
that must be added to the air so that the required humidity is achieved at any pressure
and temperature. Given that the mass of any species in a gas mixture is proportional to
the product of the molecular mass and the partial pressure and that the molecular mass
of air is usually taken to be 28.97, equation (4.3) yields:
18.016 Pw
28.97 Pa
mw
ma
0.622
Pw
Pa
(4.7)
The total air pressure P is the sum of the dry air Pa and water vapour pressures P w, so:
P
Pa
Pw
Pa
P Pw
(4.8)
Substitution of equation (4.8) into equation (4.7) and subsequent rearrangement
yields:
mw
0.622
Pw
ma
P Pw
(4.9)
The water vapour pressure, P w, can be obtained by using data from Table 4.2. The
mass of dry air per second required by a fuel cell can be found from equation (A2.10) in
Appendix 2. Note that the mass of water needed is inversely proportional to the total air
pressure P. Higher pressure systems require less added water to achieve the same
humidity. A worked example using equation (4.9) is given in Section 4.4.5.
4.4.3
Air Humidity
Previous sections have noted that the humidity of gases within the PEMFC has to be
controlled carefully to achieve the optimum level of hydration throughout the membrane
electrolyte. Fortunately, it is not difficult to derive a simple formula for the humidity of
the exit air. This is given by:
Pw
Pexit
Pw
Pexit
number of water molecules
total number of molecules
nw
nw
nO2 nrest
(4.10)
where
nw is the number of moles of water leaving the cell per second
nO2 is the number of moles of oxygen leaving the cell per second
nrest is the number of moles of the ‘non‐oxygen’ component of air per second, mainly
nitrogen
P w is the vapour pressure of the water
Pexit is the total air pressure at the exit of the fuel cell
If it is assumed that all of the product water from the cell is removed by the cathode air,
then equation (A2.17) in Appendix 2 can be used, namely:
nw
Pe
2Vc F
(4.11)
Proton‐Exchange Membrane Fuel Cells
where Pe is the power of the fuel‐cell stack and Vc is the voltage of each cell.
From equation (A2.7), the rate of use of oxygen can be expressed as:
nO2
rate of supply of O2 rate of use of O2
Hence:
nO2
1
Pe
4Vc F
(4.12)
where λ is the air stoichiometry. The exit flow rate of the inert components of air (mainly
nitrogen) will be the same as at the inlet. These components amount to 79% by volume
of air so the flow rate will be proportionately greater than the oxygen molar flow rate,
i.e., by the ratio 0.79/0.21 = 3.76, so that:
nrest
3.76
Pe
4Vc F
(4.13)
Substituting equations (4.11), (4.12) and (4.13) into equation (4.10) yields:
Pw
Pexit
Pe
2Vc F
Pe
1
4Vc F
Pe
2Vc F
2
2
1 3.76
3.76
Pe
4Vc F
(4.14)
2
1 4.76
The relationship reduces to:
Pw
0.42
Pexit
0.21
(4.15)
Thus, it is seen that the vapour pressure of water at the cathode outlet depends only on
the air stoichiometry and the air pressure at the exit Pexit. In this derivation, any water
vapour in the inlet air has been ignored, and, consequently, the formula represents the
‘worst case’ situation, with dry inlet air.
As an example, consider a fuel cell that is operating with an exit air pressure of 110 kPa,
a temperature of 70°C and an air stoichiometry of λ = 2. If the humidity of the inlet is
low, i.e., any inlet water can be ignored, then substituting these values into equation
(4.15) gives:
Pw
0.42 110
2 0.21
20.91 kPa
(4.16)
Referring to the data in Table 4.2 and using equation (4.4), the RH of the exit air is:
Pw
Psat
20.91
31.19
0.67 67%
(4.17)
97
98
Fuel Cell Systems Explained
This would be judged too dry and therefore would require attention. The humidity in
the cell could be increased by the following:
●
●
●
Lowering the cell temperature, which would increase losses.
Lowering the rate of air flow and hence λ, which would help a little, but would reduce
cathode performance.
Increasing the air (and fuel) pressure, which would require energy to run the
compressors.
Another option is to condense the water from the exit gas and use it to humidify the
inlet air. This has an obvious penalty in terms of extra equipment, weight, size and cost,
but it may be justified by the increase in performance that is possible. If the water
content of the inlet is not negligible, it can be shown that the pressure of the outlet
water vapour is given by the slightly more complex formula:
Pw
0.42
1
Pexit
0.21
(4.18)
where ψ is a coefficient with a value given by:
PWin
Pin PWin
(4.19)
where Pin is the total inlet air pressure, which will usually be slightly greater than Pexit
and PWin is the vapour pressure of water at the inlet.
Equations (4.15), (4.18) and (4.19) therefore provide a means of ensuring that there is
adequate humidity in an operating PEMFC.
4.4.4
Self‐Humidified Cells
In the example given in the previous section, the exit air from the fuel cell was too dry.
By choosing suitable operating temperatures and air flow rates, it is possible to operate
a PEMFC that has adequate internal humidification, i.e., self‐humidified. The exit air
humidity at different temperatures and air flow rates can be obtained from equation
(4.11) or equation (4.13), together with the saturated water vapour pressure taken from
Table 4.1. Examples of the exit humidity at air stoichiometries of 2 and 4 are shown in
Figure 4.18 for a cell that is operating at 100 kPa. Some selected values are also given in
Table 4.4. It can readily be seen that there is a band of operating conditions within which
an adequate level of humidification can be achieved.
As would be expected, the RH is low at high rates of air flow and falls sharply at high
temperatures. The cell will dry out, if the RH of the exit air is much less than 100% since
the data of Figure 4.18 are calculated assuming that all the water produced by the
cell evaporates. If the calculated RH is above 100%, water will condense out in the
electrodes, which will then become flooded. Consequently, in practical terms, there is
a narrow band of operating conditions imposed by the requirement to maintain an
adequate level of humidification. Providing that the temperature of the cell is maintained
below about 60°C, there will be an air flow rate that achieves an RH of around 100%.
Some of the conditions are given in Table 4.4.
Proton‐Exchange Membrane Fuel Cells
200
180
2
4
Relative humidity/ %
160
140
120
Too wet
100
80
Too dry
60
2
40
20
0
20
4
30
40
50
60
70
90
80
Temperature/°C
Figure 4.18 A graph of relative humidity versus temperature for the exit air of a PEMFC with an air
stoichiometry of 2 and 4. The entry air is assumed to be dry, and the total pressure is 100 kPa.
Table 4.3 Theoretical exit air relative humidities at selected temperatures and stoichiometries.
The inlet air is assumed to be at 20 °C and 70% relative humidity; the exit air pressure is assumed
to be at 100 kPa. The blanks in the Table are where the relative humidity is too high or too low.
Temperature (°C)
λ = 1.5
λ=2
λ=3
λ=6
20
30
194
40
λ =12
λ = 24
213
142
117
78
273
195
112
68
45
50
208
164
118
67
40
26
41
60
129
101
72
70
82
65
46
80
54
43
30
90
37
28
An important conclusion from Figure 4.18 and Table 4.4 is that at temperatures above
about 60°C (at atmospheric pressure) the RH of the exit air is below, or well below, 100%
at all reasonable values of air stoichiometry. In other words, self‐humidification can be
achieved for a cell operating at or below 60°C, but for a PMEFC operating above this
temperature, extra humidification is usually essential. This feature makes for difficulties in
choosing the optimum operating temperature for a PEMFC. The higher the temperature,
the better the performance is — mainly because of a reduction in the overpotential at
the positive electrode. Once above 60°C, however, the extra weight and cost of the
99
100
Fuel Cell Systems Explained
Water circulation
Dry air
Damp air
c
e
a
Damp hydrogen
Water circulation
Dry hydrogen
Membrane electrode
assembly (MEA)
Figure 4.19 Contra-flow of reactant gases to even the humidification throughout a cell.
additional equipment required to humidify the cell can outweigh the benefits of a
simple system as is required, for example, for a small air‐breathing fuel cell.18
One of the several ways to achieve self‐humidification is to employ a countercurrent
flow arrangement of hydrogen and air, i.e., the air and hydrogen flow in opposite
directions across the MEA, as illustrated in Figure 4.19. The water that flows through
the membrane from anode to cathode is fairly uniform across the cell, since it is driven by
the ‘electro‐osmotic drag’, which is directly proportional to the current density. The back
diffusion of water from cathode to anode decreases from the anode inlet to the anode
outlet. Even distribution of humidity is also encouraged by the use of thin electrodes
and thick GDLs to hold more water and by recycling the fuel gas. Through application
of these measures and control of the flow rate of air to match the demands of the load,
it is possible to define a band of operating parameters in which a PEMFC stack can be
self‐humidified. These requirements, however, are difficult to accomplish for systems
above a few watts capacity, and therefore many developers have opted for external
humidification as described in the following section.
4.4.5
External Humidification: Principles
It has been shown that operating temperatures of over 60°C are desirable to reduce
losses, especially the cathode activation overpotential described in Section 3.4, Chapter 3.
This objective can be achieved with external humidification and can be demonstrated by
revisiting equation (4.18). Consider, for example, a fuel cell that is operating at 90°C with
a moderate inlet pressure of 220 kPa, an exit pressure of 200 kPa and a typical air stoichiometry of 2.0. Assuming the air at the cathode inlet to be at 20°C and 70% RH, then
equation (4.16) together with values from Table 4.2 yields the following information:
●
●
●
●
Inlet water vapour pressure PWin is 1.64 kPa.
The Ψ term is 0.00751.
Vapour pressure of the water at the exit is 39.1 kPa.
Exit air humidity is 56%.
18 An air‐breathing cell is one in which the air electrode is open to the atmosphere and no forced air flow is
provided. Such devices are currently being produced for charging mobile phones.
Proton‐Exchange Membrane Fuel Cells
Under these conditions, the exit humidity is far too low and the membrane would dry
out very quickly. If, however, the inlet air is warm and damp, say, 80°C and 90% RH, the
following conditions are established:
●
●
●
●
Inlet water vapour pressure PWin is 42.65 kPa.
The Ψ term is 0.2405.
Vapour pressure of the water at the exit is 66.96 kPa.
Exit air humidity is 95%.
The amount of water that must be added to the inlet air to achieve the exit humidity
of 95% can be determined from equation (4.9). For example, if the given fuel cell that is
operating at 220 kPa, an exit pressure of 200 kPa and a typical air stoichiometry of 2.0
was a 10‐kW cell, then using equation (A2.10) from Appendix 2 the mass flow rate of
dry air (kg s−1) into the cell is given by:
a
m
3.57 10
7
2 10 103
10 103
0.65
0.011
(4.20)
The desired water vapour pressure is 42.7 kPa; see equation (4.2) and Table 4.2. The
amount of water (kg s−1) that has to be added to the air, as given by equation (4.9), is
therefore:
w
m
0.622
42.7
0.011 0.0016
220 42.7
(4.21)
The rate is approximately equivalent to 100 ml min−1. From where can this water be
obtained? Having the water as an extra input to the fuel‐cell system is obviously not
desirable so the next best option is to separate it from the cathode exit gas, since the fuel
cell generates water by virtue of the electrochemical reactions. For the above 10‐kW cell,
equation (A2.17) in Appendix 2 predicts the rate of water production to be 0.0014 kg s−1.
The total flow of water in the cathode exhaust is therefore 0.0016 + 0.0014 = 0.003 kg s−1.
Given that the water is expelled as a vapour, a condensation or separation system in the
exit path must extract rather more than half of the entrained water so that it can be
recycled for use in a humidifier. Such a supply arrangement can also ensure that the
purity of the water is maintained, but it does make the system more complicated.
Before considering some of the practicalities that are involved in humidifying the
PEMFC reactants, three factors have to be considered, as follows:
1) It is often not the case that only the air is humidified. To ensure that humidity is even
within the cell, the hydrogen fuel may be humidified as well.
2) Humidification involves evaporating water in the incoming gas. The process will
cool the gas, as the energy required to evaporate the water will come from the air.
This feature is helpful in pressurized systems because the temperature rises when air
is compressed. Consequently, the humidification process is an ideal way of lowering
the air temperature to match the inlet requirements of the stack (compressors are
discussed in Section 12.1.1, Chapter 12).
3) The quantities of water added to the air and the resulting benefits in terms of
increased humidity are all much improved by raising the operating pressure.
Conversely, lowering the pressure causes more problems. Recalculation of all the
101
102
Fuel Cell Systems Explained
values in the example used in this section, for example, but with the 10‐kW stack
operating at pressures of 140 kPa (inlet) and 120 kPa (outlet), shows that the mass
of water to be added to the inlet air stream becomes 0.003 kg s−1, and yet the exit
humidity is hopelessly inadequate at 34%. The operating pressure therefore clearly
has a major influence on humidification and is considered further in Section 4.5.
4.4.6
External Humidification: Methods
No standard method is used to humidify the reactant gases of a PEMFC stack. The
following procedures require a supply of water to be available:
1) Bubbling the gases through water at a controlled temperature. The process is known
as ‘sparging’, and it is generally assumed that the dew point of the humidified air is
the same as the temperature of the water it has bubbled through, which makes
control straightforward. The procedure is suitable for conducting experimental and
test work in the laboratory, but is not a preferred method for practical systems.
2) Direct injection of water into the feed gas(es) as a spray. This technique has the
advantage that the cold water will cool the gas, an action that will be necessary if the
gas has been compressed or if it is hot through being produced by reforming some
other fuel. The method requires a pump to pressurize the water together with a valve
to open/close the injector. It is therefore fairly expensive in terms of equipment and
parasitic consumption of energy. Nevertheless, the practice is based on a mature
technology and is widely used, especially for larger fuel‐cell systems.
3) Direct injection as a fine water spray through a metal foam. This approach has the
advantage that only a pump is required to move the water — the water is injected
passively.
4) Humidification of the GDL via a series of wicks. The wicks dip into the water and
draw it directly to the GDL. The system is somewhat self‐regulating, as no water will be
extracted if the wicks are saturated. Unfortunately, the method creates a gas‐sealing
issue, namely, the wick offers an easy exit route for reactant gases. The possibility of
cooling the incoming air is also lost.
5) Direct injection of liquid water into the fuel cell. Normally, such action would lead to
flooding of the electrode and consequent failure of the cell. The technique, however,
is combined with a bipolar plate that has an ‘interdigitated’ flow‐field design (see
Section 4.6.4), which forces the reactant gases to blow the water through the cell and
over the entire electrode. The ‘flow‐field’ cut into the bipolar plate is like a maze with
no exit, as illustrated in Figure 4.20. The gas is forced under the bipolar plate and into
the electrode and thereby drives the water with it. If the flow-field is well designed, a
uniform distribution of water will be obtained all over the electrode. Good results
have been reported for direct water injection; although there are concerns that it
may degrade the electrodes over time. In addition, cooling the incoming air is not
possible with this method.
Alternatively, the following three methods enable water to be recycled from the
cathode exhaust gas:
1) Water in the cathode exit gas is used without condensation to liquid. The practice
involves the use of a rotating wheel that contains water‐absorbing or desiccant material.
The device is usually called an ‘enthalpy wheel’ and is applied in other technologies,
Proton‐Exchange Membrane Fuel Cells
Out
In
Top view
Bipolar plate
Gas and water driven
through electrode
Electrolyte
Side view, enlarged
Figure 4.20 Diagrams to show the principle of humidification using interdigitated flow-fields.
(Source: After Wood, DL, Yi, JS and Nguyen, TV, 1998, Effect of direct liquid water injection and
interdigitated flow-field on the performance of proton exchange membrane fuel cells,
Electrochimica Acta, vol. 43(24), pp. 3795–3809. Reproduced with permission of Elsevier.)
such as air‐conditioning systems. Water in the exhaust gas is absorbed on the material,
which then rotates so that it is introduced into the path of the dry cathode inlet. The
process is continuous — it constantly delivers water from exit to inlet gases and
transfers heat from the exhaust stream to the inlet stream. The method suffers from
the disadvantages of being fairly bulky and requiring power and control system for
its operation.
2) A more sophisticated system, first disclosed by the Paul Scherrer Institute in
Switzerland, also uses the exit water without condensation. In this case, a membrane
is placed between the cathode exit and the cathode inlet streams. Water vapour in
the exit stream condenses on the membrane and then passes through it to the dry
inlet side. The membrane can be the same material as the PEMFC electrolyte, and
some manufacturers have employed this technique for every cell within a stack.
3) Many developers have sought to modify the PEMFC membrane to enhance water
retention. One approach is to generate water in situ. The membrane is modified, not
only to retain water but also to produce water. Retention is enabled by impregnating
the electrolyte with particles of silica (SiO2) and/or titania (TiO2), together with
nanocrystals of platinum. If the membrane is sufficiently thin, the platinum catalyses
the reaction between the incoming hydrogen and oxygen to generate water. The
reaction of course uses up some valuable hydrogen gas, but it is claimed that the
improved performance of the electrolyte justifies the parasitic loss of fuel.
103
104
Fuel Cell Systems Explained
4.5
Cooling and Air Supply
4.5.1
Cooling with Cathode Air Supply
An electrical generation efficiency of around 50% may be achieved when converting the
chemical energy in hydrogen into electricity in a PEMFC, i.e., heat and electricity are
produced in approximately equal amounts. Removal of the heat is essential as the cell
has to be cooled to maintain the required operating temperature. If the product water is
evaporated within the cell, the heat produced is given by (see Section A2.6, Appendix 2):
Heating rate
Pe
1.25
1
Vc
(A2.21)
The way this heat is removed depends greatly on the size of the fuel cell or stack. For
fuel cells below about 100 W, it is possible to use naturally flowing air to cool the cell
and to evaporate the water produced without recourse to a fan. Similar convective
cooling can be applied to stacks that have a fairly open structure with a spacing of
between 5 and 10 mm per cell. With a more compact design of fuel cell, small ‘fans’ can
be employed to blow excess air through the cell cathodes, though a large proportion of
the heat will still be lost through natural convection and radiation. For small systems,
such an air fan only imposes a minor parasitic loss of power on the system, namely,
about 1%, for a well‐designed system. For systems with stacks producing more than
about 100 W, proportionately less heat is lost by natural convection and radiation from
and around the external surfaces of the cells. Larger systems therefore require forced
cooling in addition to that provided by the cathode air to maintain the necessary low
operating temperature.
4.5.2
Separate Reactant and Cooling Air
The need to separate the reactant air and the cooling air for anything but the smallest of
PEMFCs can be demonstrated by working through a specific example where the
reactant gas and the cooling gas are combined. Consider, therefore, a fuel cell of power
Pe watts running at 50°C with each cell in the stack at an average voltage of 0.6 V. Suppose
that cooling air enters the cell at 20°C and leaves at 50°C. (In practice the temperature
change will probably not be so great, but it is instructive to take the best possible case for
the present.) Assume also that only 40% of the heat generated by the fuel cell is removed
by the air — the rest is radiated or naturally lost by convection from the outer surfaces.
The rate of removal of heat by air of specific heat capacity CP flowing at a rate of
kg s−1, and subject to a temperature change ΔT will be the same as the heat produced
m
according to equation (A2.21). It follows that:
0.4 Pe
1.25
1
Vc
P T
mC
(4.22)
On substituting known values, i.e., CP = 1004 J kg−1 K−1, ΔT = 30 K and Vc = 0.6 V and
then rearranging, the following equation is obtained for the flow rate of the cooling air:
1.4 10
m
5
Pe
(4.23)
Proton‐Exchange Membrane Fuel Cells
In Section A2.2, Appendix 2, it is shown that the flow rate of reactant air is:
3.58 10
m
7
Pe
Vc
(4.24)
If the reactant air and the cooling air are one and the same, then these two flow rates
are equal. Therefore, combining equation (4.23) and equation (4.24), cancelling Pe,
substituting Vc = 0.6 V and solving for λ yields:
14 0.6
0.357
24
(4.25)
Reference to Table 4.2 will show that at 50°C, this air stoichiometry produces an exit air
humidity of 27%, i.e., dryer than the Sahara desert! The data in Table 4.2 assume that the
entry air has a humidity of 70%. Consequently, the RH is decreasing as the air goes
through the cell, and this will promote rapid drying out of the PEM.
If the assumptions made at the beginning of this section are made more realistic, that
is, more heat has to be taken out by the air because the cell efficiency is lower, and the
allowable temperature rise in the cell is also lower to maintain adequate humidity,
then the situation becomes even worse. The only way to reduce λ, which should be
somewhere between 3 and 6 at 50°C in order to prevent the cell from drying out, is to
decrease the rate of the air flowing over the electrodes and have a separate cooling
system. Such action is necessary when more than about 25% of the heat generated by a
fuel cell has to be removed by a cooling fluid. In practice, this applies to cells of about
100 W in size. Fuel cells of greater output will generally need a separate forced supply of
reactant air and a cooling system.
The usual way to cool cells with outputs from about 100 W to 1 kW is to make extra
channels in the bipolar plates through which cooling air can be blown, as shown in
Figure 4.21. An alternative approach is to add separate cooling plates for the passage
of air. A commercial air‐cooled stack is shown in Figure 4.22. For systems larger than
1–5 kW, air is no longer adequate and water cooling is preferred.
4.5.3 Water Cooling
Cooling with air is the simplest option for PEMFC stacks and is generally adopted for
those below about 2 kW. Above about 5 kW, the common preference is to employ water
cooling given its advantage of not requiring large channels for the cooling medium — 1 kg
of water can be pumped through a much smaller channel than 1 kg of air, and the
cooling effect of water is much greater. Water‐cooled stacks can therefore be more
compact for a given kW size and, additionally, bring the benefit of being able to employ
the heat generated by the fuel cell, e.g., in a domestic CHP system. With an air‐cooled
system, heat is lost to the atmosphere, whereas heat from a water‐cooled system can be
put to practical use.
The method of water cooling a fuel cell is essentially the same as for air in Figure 4.21,
except that water is pumped through the cooling channels. In practice, such channels
are not always necessary or provided at every bipolar plate. The following section
considers cooling in more detail for manufactured systems.
105
106
Fuel Cell Systems Explained
MEA with sealing gasket
on each side
Cooling air
through these
channels
Reactant air
feed channels
Hydrogen feed
channels
Figure 4.21 Three‐cell PEMFC stack with bipolar plates with separate reactant and cooling air
channels.
Horizon 5000 W Air-breathing fuel-cell stack
Number of cells................................................................120
Rated performance...............................................72 V@70 A
Reactants...................................................Hydrogen and Air
Ambient temperature................................5 – 30°C (41 – 86°F)
Max stack temperature....................................65°C (149°F)
Hydrogen pressure.........................................0.45 – 0.55 Bar
Humidification...............................................Self-humidified
Cooling.......................................Air (integrated cooling fan)
Weight (with fan and casing).............................30 kg (±200 g)
Controller weight...........................................2500 g (±100 g)
Stack size........350 x 212 x 650 mm (13.8 × 8.3 × 25.6 in)
Flow rate at max output............................................65 L/min
Start-up time.............................≤30 s(ambient temperature)
Efficiency of system.............................................40% @72 V
Figure 4.22 Example of a commercial air‐cooled stack manufactured by Horizon Fuel Cells. (Source:
Reproduced with permission of Horizon Fuel Cells.)
Proton‐Exchange Membrane Fuel Cells
4.6
Stack Construction Methods
4.6.1
Introduction
Most PEMFC stacks are constructed along the general lines of multiple cells connected in
series with bipolar plates, as outlined in Section 1.3, Chapter 1 and illustrated in Figure 1.7.
The bipolar plate has to collect and conduct the current from the anode of one cell to the
cathode of the next, while distributing the fuel gas over the surface of the anode, and the
oxygen/air over the surface of the cathode. Furthermore, the plate often has to carry a cooling
fluid though the stack and keep all the reactant gases and cooling fluids apart. The
distribution of the reactant gases over the electrodes is achieved with a ‘flow-field’ formed
into the surface of the plate. The flow field usually has a fairly complex serpentine pattern.
Bipolar plates contribute a high proportion of the cost of a PEMFC stack and have to
satisfy several requirements, namely:
●
●
●
●
●
●
Good electrical conductivity (>100 S cm−1).
High thermal conductivity — this should exceed 20 W m−1 K−1 for normal integrated
cooling fluids or must exceed 100 W m−1 K−1 if heat is to be removed only from the
edge of the plate.
High resistance to chemical attack and corrosion.
High mechanical stability, especially under compression (flexural strength >25 MPa).
Low gas permeability (<10−5 Pa L s−1 cm−2).
Low density — to minimize both weight and volume of the stack.
The methods of forming the plates, as well as the materials from which they are fabricated,
vary considerably. Similar to humidification, which was considered in the previous
section, there is no single method or material that is optimum for every application.
Before examining the materials and the manufacturing methods, it should be
understood that in most cases the plate is made in two halves. Cooling channels are cut
on the back of one of the half‐plates. Whereas this design simplifies the incorporation
of cooling within the assembled plates, a significant electrical resistance can arise
between the two half‐plates when they are pressed together. The latter feature is
especially a problem for carbon plates. For metal plates, there are several methods (e.g.,
welding, diffusion bonding) that can be used to join half‐plates. The first bipolar plates
were machined from sheets of graphitic carbon, and this material remains a good choice
for stationary applications of fuel cells where longevity is required and compactness is
less of a priority. For service in vehicles, stainless steel tends to be favoured given that it
can be formed into very thin sheets by means of processes comparable with those
practised in the automotive industry. Metal bipolar plates usually require a surface
coating to protect against both chemical attack and corrosion.
Nearly all PEMFC stacks incorporate either carbon or metal bipolar plates, as described
in the following two subsections. Other types of cell connection and topologies have been
investigated for some smaller systems, and examples are given at the end of this section.
4.6.2
Carbon Bipolar Plates
Graphite has high electrical and thermal conductivity. It also has a very low density, i.e.,
less than for any metal that might be considered suitable for a bipolar plate, as well as
good resistance to chemical attack. Consequently, the earliest PEMFCs had graphite
107
108
Fuel Cell Systems Explained
bipolar plates into which the flow‐field channels were machined. Graphite does,
however, have the following three disadvantages:
1) The plates must have a thickness of few mm to provide the mechanical integrity
required for machining and handling, even though the latter activity may be
automatic; cutting of graphite is time‐consuming with an expensive milling machine.
2) Graphite is brittle, so the resulting cell demands careful handling.
3) Graphite is quite porous, so plates have to be coated and made sufficiently thick to
ensure that the reactant gases are separated. Therefore, although the density of
graphite is low, the final bipolar plate may not be particularly light in weight.
These disadvantages are addressed by using composite materials that combine graphite
powder with a polymer binder. Such materials have also been used for PAFCs. Most
state‐of‐the‐art carbon bipolar plates for the PEMFC are made from a composite that
consists of a high loading of conductive carbon (e.g., graphite, carbon black or
nanotubes)19 with a commercial thermoplastic polymer binder (e.g., polyethylene,
polypropylene or polyphenylene sulfide) or thermosetting resins (phenolic or epoxy
resins). Carbon fibres are often added to strengthen the finished product.
Since the physical properties of the composite carbon materials are largely determined
by the polymer binder, bipolar plates can be formed either by compression moulding
or by injection moulding. The former procedure requires the use of a thermoplastic
polymer and is usually the chosen method for limited quantities of small stacks. The
mould has a top and bottom part. Graphite powder mixed with polymer is spread over
the lower part of the mould. The top is lowered in place and pressure applied. The
temperature is then raised above the melting point of the polymer so that the materials
mix and flow to fill the mould. After cooling, the product can be released. High carbon
content can be achieved with compression moulding. The process is simple but unfortunately slow — production times of 15 min are typically needed for each half‐plate.
Variations of the compression moulding technique have been published. For example,
the porous material that normally results from compression moulding can be made
less permeable by applying a coating of solid carbon on the back of the plate. This
modification is realized by chemical vapour infiltration, which is a standard technique
easily adapted to mass production. In another procedure, carbon black is taken as the
raw material for compression moulding, and the resulting plate is heated to a high
temperature (above 2500°C) to cause graphitization. Although this method can improve
the electrical conductivity, it can also give rise to warped and brittle sheets.
Injection moulding is an attractive proposition for larger and more complex components
and involves the participation of a thermosetting polymer. The process is, however, very
demanding in that the composite has to be sufficiently fluid to flow into the mould, while
at the same time have an adequate carbon loading to achieve good electrical conductivity.
Thermosetting polymers also have very different properties to their thermoplastic
counterparts. Consequently, composites of the former are injected as a powder into a
mould at a temperature below the melting point of the thermosetting polymer, whereas
the latter composites are heated above the melting point of the thermoplastic polymer
and injected into a mould that is maintained at a lower temperature.
19 Natural graphite flakes have also been used; these are processed to produce a continuous foil that can
then be made gas‐tight by incorporating a polymer binder.
Proton‐Exchange Membrane Fuel Cells
Whereas thermosetting polymers may allow higher loadings of carbon, a time‐
consuming curing step is required after the moulding is complete. With both techniques,
the surface of the resulting injection‐moulded plate has to be cleaned, e.g., with an abrasive,
to remove any film of polymer that would limit the electrical contact with the GDL.
Although many of the details are proprietary, several companies are now manufacturing bipolar plates with thermoplastic‐bonded graphite structures. These alternatives
offer a route to reduced costs through mass production.
4.6.3
Metal Bipolar Plates
Metals have advantages over carbon in that they are good conductors of heat and
electricity, can be machined easily and are not porous. The main disadvantages are that
they have higher density and are prone to corrosion — the hot oxygen and water vapour
inside a PEMFC is quite a corrosive environment. In addition, there is sometimes a
problem of acid leaching out of the MEAs, and therefore it is common practice to coat
metal bipolar plates with a corrosion‐resistant material. Plates made of stainless steel,
titanium, aluminium and several alloys have been tested with separate coatings of a
conductive carbon–polymer material, a transition metal that forms a passive oxide
layer (e.g., molybdenum, vanadium or niobium), or a noble metal (e.g., gold).
Metal sheets can be machined in similar fashion to graphite sheet, but the process is
expensive. Nevertheless, the fact that metals are not as brittle and porous as carbon
means that thinner plates can be made. If the plates are sufficiently thin, flow fields can
be made by stamping the patterns onto the metal plates. Although stamping is widely
practised industrially, it is difficult to achieve extremely narrow channels (with depth
and width measured in mm) by stamping the metal.
An alternative approach is to use perforated metal or metal foam to form the flow
fields. One such method has been described by Murphy et al.20 who have chosen
titanium as the metal. The electrical conductivity of titanium is relatively low compared
with metals such as copper that exhibit good conductivity — nevertheless, it is some
thirty times greater than graphite. As a material for a bipolar plate, titanium can also
be made sufficiently corrosion‐resistant by coating with titanium nitride. Such an
electrically conductive coating can be applied inexpensively on a large scale. The
method adopted by Murphy for creating a bipolar plate is to employ two sheets of metal
foam with a thin layer of solid metal between them. The concept is shown in Figure 4.23.
The pores or voids in the foam sheets serve as pathways for the diffusion of gas to the
electrodes. The reactant gases are fed to the edges of the foam sheets via porous plastic
gaskets that are sealed to the periphery of the plate.
Metal foam can also be used to effect the cooling of a stack. To do this, one sheet of
the metal foam is placed between two sheets of solid (but thin) metal sheets. Water is
passed through the metal foam to carry away the heat. The method offers the advantage
of using readily available materials (metal foam sheet is made for other applications) to
make fuel‐cell components that are thin, lightweight, highly conductive and serve to
separate the reacting gases. Furthermore, the only manufacturing processes involved in
employing foams are the cutting and moulding of the plastic edge seals. There are many
20 Murphy, OJ, Cisar, A and Clarke, E, 1998, Low cost light weight high power density PEM fuel‐cell stack,
Electrochim. Acta, vol. 43(24), pp. 3829–3840.
109
110
Fuel Cell Systems Explained
Metal sheet
Edge seals with holes
to feed gas into metal
foam
Complete bipolar
plate
Metal foam
Metal foam
Figure 4.23 Diagram showing bipolar plate construction from metal foam. (Source: After Murphy, OJ,
Cisar, A and Clarke, E, 1998, Low cost light weight high power density PEM fuel‐cell stack, Electrochim
Acta, vol. 43(24), pp. 3829–3840. Reproduced with permission of Elsevier.)
ways in which this basic idea can be adapted, but in each case the foam needs to be
coated to protect against corrosion.
In summary, coated metal or metal foam bipolar plates have been successfully applied
to vehicle fuel‐cell stacks where corrosion stability is not as demanding as for stationary
power generation. For vehicles, a minimum stack lifetime of 2 000 h is required, whereas
for stationary systems 40 000 h is expected and is now routinely achieved with carbon
bipolar plates.
The rapid growth in the application of 3‐D printing in recent years has not been
ignored by fuel‐cell researchers. Many research groups have reported the fabrication of
both carbon composite and metal bipolar plates using 3‐D printing methods, which are
particularly promising for printing the flow‐field patterns described in Section 4.6.4.
Unfortunately, current methods lack the precision for fabricating large numbers of
components and have generally been employed only for creating prototype stacks.21
4.6.4
Flow‐Field Patterns
In the bipolar plate illustrated in Figure 1.12, Chapter 1, the reactant gases are fed over
the electrode in a simple pattern of parallel grooves. The other three basic types of flow‐
field pattern chosen for bipolar plates are the ‘pin (or grid)’, ‘interdigitated’ and
‘serpentine’ varieties; the four alternatives are illustrated in Figure 4.24a–d. The design
of flow field is influenced both by the plate material itself and by the adjoining GDL. The
aim with any design of flow-field is to ensure that humidity is balanced throughout the cell
and that gases can flow readily to and from each GDL. It is also desirable to minimize
21 Gould, BD, Rodgers, JA, Schuette, M, Bethune, K, Louis, S, Rocheleau, R and Swider‐Lyonsa, K, 2015,
Performance and limitations of 3D‐printed bipolar plates in fuel cells, ECS Journal of Solid State Science and
Technology, vol. 4(4), pp. 3063–3068.
Proton‐Exchange Membrane Fuel Cells
(a)
(b)
(c)
(d)
(e)
(f)
(g)
(h)
Figure 4.24 Main flow‐field patterns used in PEMFC bipolar plates: (a) pin or grid type, (b) parallel
channels, (c) interdigitated, (d) single serpentine channel (e–h) oxygen concentration in the cathode
GDL for each of the flow‐field patterns (schematic). Red, high oxygen concentration; blue, low oxygen
concentration. (Source: From Heinzel, A, Mahlendorf, F and Jansen, C, 2009, Bipolar plates, in Garche, J,
Dyer, C, Moseley, P, Ogum, Z, Rand DAJ and Scrosati B (eds.), Encyclopedia of Electrochemical Power
Sources, pp. 810–816, vol. 2, Elsevier, Amsterdam.)
the pressure drop through the flow-field. Inevitably, some compromise has to be reached.
As has been remarked earlier, the configuration of flows on either side of the MEA, i.e.,
co‐flow, crossflow or counterflow, also has an important bearing on cell performance.
The parallel‐channel arrangement, Figure 4.24b, can be used when there is little
likelihood of the formation of water droplets, which otherwise could accumulate, blocking
some the channels and thereby cause poor distribution of current throughout the cell
and stack. The development of droplets can be minimized by ensuring that flows are
large, e.g., in the case of the fuel flow channel, by recycling of the fuel gas.
The pin flow-field, Figure 4.24a, is also best suited to applications that require high
reactant flows and low utilization of fuel and oxygen. In this arrangement, gases can
swirl all over the face of the electrode. Unfortunately, any slight perturbation in the flow
can lead to a path of least resistance through the flow field. Such behaviour can engender
uneven distribution of reactants, especially on the oxygen side.
The interdigitated flow field, Figure 4.24c, consists of a large number of dead‐ended
channels. In this arrangement, the gases are forced to flow through the GDL, which
therefore has to have adequate porosity and hydrophobicity. Since the efficacy of these
two parameters may decline as the GDL ages, this pattern is not generally favoured.
The serpentine pattern is preferred by most PEMFC manufacturers. It offers a good
compromise between the issues of pressure drop and water removal. The large number
of turns in the flow path means that pressure drop is compromised, and this is
overcome usually by employing multiple parallel serpentines rather than the single
example shown in Figure 4.24d.
111
112
Fuel Cell Systems Explained
The flow of fluids within the channels of a bipolar plate is difficult to measure directly.
Accordingly, considerable work has been undertaken to model flows with the assistance
of fluid‐flow finite‐element methods. Many simulations of gas flows within PEMFC
stacks have been published in the literature. For example, the concentration of oxygen
at the cathode GDL predicted for the four different flow-fields is presented in
Figure 4.24e–f. Several research teams have also developed the technology to measure
the current density distribution within an operating fuel‐cell stack. This is commonly
done by inserting a thin sensor plate between adjacent cells. The plate usually has a
large number (typically over 100) of individual segments that are electrically insulated
from each other, but can be connected by fine wires to a measuring device. Each
segment imposes a small electrical resistance at regular intervals between adjacent
cells, and measurement of the voltage developed across the resistances enables
determination of the local current density. Data obtained from working stacks not only
help to validate models but also assist in optimizing the design of flow-fields and other
cell components.
4.6.5
Other Topologies
The construction of fuel‐cell stacks through the application of bipolar plates gives very
good electrical connection between one cell and the next. On the other hand, the use
of bipolar plates necessarily means that there are many joints with the potential for
causing leakage of both reactant gas and cooling fluid. The supply of reactant gas to
each and every positive electrode has to be kept separate from that to each and every
negative electrode. The entire edge of each anode and cathode is also respectively
subject to leakage. Good quality control in the production of components minimizes
the risk of leakage, but necessarily means high fabrication costs.
In situations where a fuel cell is operated at a fairly low current density, it is often
useful to compromise the electrical resistance of the cell interconnects in the interest of
simpler and cheaper manufacturing methods. This option is available because the
flexibility and ease of handling of the MEAs used in PEMFCs allows for different types
of construction other than the traditional bipolar plate. One such design is the system
of three cells shown in Figure 4.25. The main body of the unit (depicted in light grey)
would normally be made of plastic material. There is only one chamber containing air
and only one containing hydrogen. By passing a metal connector strip through the
reactant gas separator, the cells are joined in series with the edge of one cathode
connected to the edge of the next anode. For even less chance of leaks, the connection
MEA with gas-diffusion layers
above and below. Perforated metal
current collectors are attached to
the gas diffusion layers
c
c
Air
c
a
a
Hydrogen
a
+
Figure 4.25 Method of connecting fuel cells in series to simplify reactant gas supply.
Proton‐Exchange Membrane Fuel Cells
could be made externally but this would increase the current path. In this design, the
potential for leaks is greatly reduced as the only seals are those around the edges of the
MEAs, and the challenge of maintaining uniform humidification of the cells is simplified since there is fairly free circulation of reactant gases in the cell. In practice, however,
the arrangement is not compact and is therefore only suitable for low‐power systems.
Small PEMFCs can also employ a cylindrical stack design. For example, a microtubular
design was designed by NASA for small electronics applications. In most cases,
however, the fuel cell is preferably assembled on a planar substrate as a two‐dimensional
device with edge connections between cells. Printed‐circuit and microelectromechanical
system (MEMS) technologies are preferred as these are amenable to well‐established
mass-production processes. The MEMS approach has given rise to designs such as
those illustrated in Figure 4.26. Such small systems are beginning to enter the market
after many years of research and development.22 For example, Ultracell has an exclusive
licence with Lawrence Livermore National Laboratory for a micro‐fuel‐cell system
targeted principally at military applications.
Subtle variations of the traditional planar fuel cell have been investigated in an attempt
to increase the power density of stacks. Intelligent Energy has pursued the design of an
air‐breathing stack in which each cell is constructed from a stainless steel anode
current-collector with flow-fields for the hydrogen on top of which sits the MEA, and
(a)
(b)
Air-breathing holes
Current collector
Computational domain
Figure 4.26 Examples of MEMS‐based air‐breathing PEMFCs. (a) Six cells sealed with epoxy resin
(Si and glass wafers, 1.2 cm × 1.2 cm single‐cell active area, ~4 cm3, >140 mW cm−2 peak power with
H2)23. (b) PCB‐based air‐breathing PEMFC with other integrated components24.
22 Pichonat, T and Gauthier‐Manuel, B, 2006, Recent Developments in MEMS‐based micro fuel cells,
DTIP, Stresa, Lago Maggiore, Italy. TIMA Editions 6p. <hal‐00189312>. Available https://hal.archives‐
overtes.fr/hal‐00189312 (accessed on 15 August 2017).
23 Reprinted from Zhang, XG, Wang, T, Zheng, D, Zhang, J, Zhang, Y, Zhu, L, Chen, C, Yan, J, Liu, HH,
Lou, YW, Li, XX and Xia, BJ, 2007, Design, fabrication and performance characterization of a miniature
PEMFC stack based on MEMS technology, International Journal of Electrochemical Science, vol. 2,
pp. 618–626.
24 Reprinted from Hwang, JJ and Chao, CH, 2007, Species‐electrochemical transports in a free‐breathing
cathode of a PCB‐based fuel cell, Electrochimica Acta, vol. 52, pp. 1942–1950.
113
114
Fuel Cell Systems Explained
Metallic, porous, conductive
and water-retaining cathode
current collector
More cells
Channels through
which hydrogen
flows
Insulator
Stainless steel
cell interconnector
Stainless steel
anode current
collector and
flow-field for
hydrogen
MEA with gasdiffusion layers
Open structure
allows free
circulation of
cooling and
reactant air
More cells
Figure 4.27 Structure PEMFC demonstrated by Intelligent Energy.
then a porous metal current collector sits on top of the cathode. Patented and proprietary
techniques are employed to fabricate the cathode current-collector from sintered,
stainless steel powder of carefully graded size. The result is a material that is metallic,
corrosion resistant, porous, strong, conductive and water retaining. A fuel‐cell stack is
assembled by placing the self‐contained cells one on top of the other; a simple piece of
folded stainless steel connects the anode of one cell to the cathode of the next. The
arrangement is shown in Figure 4.27. Hydrogen is piped via thin plastic tubing to each
anode. The open structure of the cell allows for free circulation of air, though this may
be fan assisted.
4.6.6
Mixed Reactant Cells
In all conventional fuel cells, fuel and oxidant are supplied as separate streams to the
anode and cathode, respectively. By comparison, in a mixed reactant fuel cell (MRFC) a
mixture of fuel and oxidant flows through the cell as a single stream. The concept first
appeared in the literature in the 1960s and is attractive in that there is no requirement
for the gas‐tight seals, which are necessary for manifolds and for separating air and fuel
systems in conventional stacks. By avoiding the cost and weight of bipolar plates in the
design, significant cost reduction should also be possible. Similarly, some simplification
of the balance‐of‐plant is to be expected. An MRFC requires the following properties:
●
The cathode catalyst should support the reduction of oxygen and not the oxidation of
fuel, i.e., mixed potentials should not be possible.
Proton‐Exchange Membrane Fuel Cells
●
●
The cell should operate at a low enough temperature to avoid the spontaneous thermochemical reaction between fuel and oxidation that may occur in the bulk reaction
mixture or on catalyst surfaces.
The electrode structures, i.e., the GDLs, should enable the fuel and oxidation to reach
the anode and cathode catalyst layers, respectively, by controlling the diffusion of
the species. Alternatively, the electrode catalysts should have sufficiently different
reaction kinetics to ensure that the fuel oxidation reaction and the ORR are separated.
In reality, none of these three properties can be 100% effective, with the result that the
cell voltage and energy efficiency of the MRFC is compromised. Rather, the issue is
whether such deficiencies are offset by potentially lower capital costs and higher power
densities that, in some applications, may favour MRFCs over conventional systems.
In principle, several different types of MRFC could be constructed according to the
type of electrolyte and cell reactions. This includes cells based on PEMFC, AFC and
SOFC materials. One of the first MRFCs to incorporate PEMFC materials was reported
in 2002.25 There followed in 2004 a direct methanol MRFC with a Pt–Ru–C anode
catalyst and a Ru–Se–C cathode catalyst, from which power densities of approximately
50 and 20 mW cm–2 could be obtained at 90°C with oxygen and air fed cathodes,
respectively.26 The mixed reactant DMFCs did not exhibit parasitic direct reaction of
methanol with oxygen.
4.7
Operating Pressure
4.7.1 Technical Issues
Although small PEMFC stacks are operated at normal air pressure; larger stacks of
10 kW or more are sometimes run at higher pressures. Increasing the operating
temperature increases the cell voltage, but as mentioned in Section 4.2.1, the PSFA
membranes need to remain hydrated. At atmospheric pressure this limits the operating
temperature to about 80°C. Raising the pressure enables the temperature to be increased.
Energy is consumed, however, in compressing the fuel and air, and may not be recovered
from the fuel‐cell exhaust streams.
The simplest type of pressurized PEMFC system is that in which the hydrogen is
supplied from a high‐pressure cylinder. Such as system, as employed for example by
Hydrogenics, is shown in Figure 4.28. Only the air has to be compressed. The hydrogen
gas is fed from a pressurized storage container to the fuel‐cell anodes. The fuel side
of the stack is ‘dead-ended’, i.e., there is no exhaust stream for the fuel gas; it is all
consumed by the cell.27 The compressor for the air has to be driven by an electric
25 Priestnall, MA, Kotzeva, VP, Fish, DJ, and Nilsson, EM, 2002, Compact mixed‐reactant fuel cells, Journal
of Power Sources, vol. 106, pp. 21–30.
26 Scott, K, Shukla, AK, Jackson, CL, Meuleman, WRA, 2004, A mixed‐reactants solid‐polymer‐electrolyte
direct methanol fuel cell, Journal of Power Sources, vol. 126(1–2), pp. 67–75.
27 Dead‐ended systems usually release a very small amount of the fuel gas from the system at regular
intervals to avoid build‐up of contaminants in the negative electrode. It is also common to recycle some of
the fuel gas back to the inlet of the stack (shown by dotted lines in Figure 4.28). Again, this helps to purge
contaminants and maintain uniformity of humidity throughout the negative electrode.
115
116
Fuel Cell Systems Explained
Anode pressure
regulator
Hydrogen recycle
compressor
Hydrogen
purge
PEM fuel cell
Anode
Compressed
hydrogen
storage
Motor
Electrolyte
Cooler/
humidifier
Cathode
Exhaust air
and water
vapour
Air
compressor
Air intake
Figure 4.28 Schematic representation of a simple PEMFC system, as employed in Hydrogenics
fuel-cell modules. (Source: Reproduced with permission of Hydrogenics.)
motor, which of course uses up some of the valuable electricity generated by the fuel
cell. Note that for the system outlined in Figure 4.28, the pressure of the hydrogen at
the anode side of the stack can be controlled as a function of the pressure on the
cathode side that, in turn, is determined by the power delivered to the air compressor.
Consequently, the differential pressure developed between the two sides of the stack
can be maintained at a constant low level to minimize the risk of gas crossover. In a
worked example in Appendix 3, it is shown that the typical power consumption by the
air compressor will be about 20% of the fuel‐cell power for a 100‐kW system.
Compression also raises the temperature of the air so that cooling may be necessary
before its entry to a PEMFC, so‐called ‘intercoolers’ operate similarly in internal
combustion engines.
When the hydrogen fuel is derived from other hydrocarbons, such as methane, the
situation is much more complex. Depending on the design of the reformer (described
more fully in Chapter 10), the fuel gas is likely to contain other components in
addition to hydrogen. In such a situation, running the fuel cell ‘dead-ended’ is therefore
not an option, and the exhaust gas stream from the anode may contain a significant
amount of unconverted hydrogen. Clearly, this fuel cannot be wasted, and the role of
the fuel‐cell system designer is to make sure that any energy in the exhaust stream is
utilized effectively. For example, unreacted hydrogen may be burned and the energy
release directed to compression of the fuel gas, or it may provide heat for the endothermic
reforming reaction.
Proton‐Exchange Membrane Fuel Cells
4.7.2
4.7.2.1
Benefits of High Operating Pressures
Current
The increase in power that results from operating a PEMFC at elevated pressure is mainly
the result of the reduction in the cathode activation overpotential, as discussed in
Section 3.4, Chapter 3. The increased pressure raises the exchange‐current density, which
in turn causes an increase in the open‐circuit voltage (OCV) of the cell, as shown in
Figure 3.4, Chapter 3. Note that, however, there is sometimes a reduction in the masstransport losses, with the result that the cell voltage begins to decline at a high current density.
The influence of pressure on cell performance can be appreciated from the graph of
voltage against current given in Figure 4.29. In simple terms, for most values of current
density, the voltage is raised by a fixed value. Although not shown by the graph, this
voltage ‘boost’, ΔV, is proportional to the logarithm of the pressure rise. The feature is
observed experimentally and has a theoretical basis. In Section 2.5.4, Chapter 2, it was
noted that the rise in OCV due to the change in Gibbs free energy can be expressed as:
RT
P
ln 2
4F
P1
V
(2.45)
As given by equation (3.8) in Chapter 3, the activation overpotential is related to the
exchange‐current by a logarithmic function. Therefore, to a first approximation, it follows
that an increase in pressure from P1 to P2 will promote an increase or gain in voltage, i.e.,
V gain
C ln
P2
P1
(4.26)
Cell voltage
Higher pressure P2
The voltage change is
fairly constant at most
currents
Normal atmospheric
pressure P1
Higher current
before masstransport losses
become important
Current
Figure 4.29 Effect of increasing pressure on the voltage versus current relationship for a typical
fuel cell.
117
118
Fuel Cell Systems Explained
where C is a parameter with a value that depends not only on how the exchange‐current
density, io, is affected by pressure, but also on the temperature. Various values for C of
between 0.03 and 0.10 V are quoted in the literature; this parameter is also influenced
by the level of cell humidification.
The simple system shown in Figure 4.28 is a useful basis for reaching an understanding
of the cost benefit offered by pressurization. For this system, the advantage lies in the
greater electrical power obtained from the fuel cell. The increase in voltage for each cell
in the stack, ΔVgain, is expressed by equation (4.26). To quantify the power gain, consider
a current of I amps flowing through a stack of n cells. The increase in power (watts) is
then given by:
Power gain C ln
P2
I n
P1
(4.27)
Some of the power produced by the fuel‐cell stack is required to drive the air
compressor. As shown later by equation (12.10), Chapter 12, an equation can be written
for the power consumed in terms of the compressor efficiency, ηc, the entry temperature
of the air T1 and the pressure ratio P2: P1, namely:
1
Compressor power C P
T1
C
P2
P1
1 m
(12.10)
is the flow rate of the air, in kg s−1. This is the power required by the
In this equation, m
compressor’s rotor. If the efficiency of the motor and drive system is expressed as ηm,
then the electrical power required by the compressor will be greater by a factor of 1/ηm.
Therefore, the electrical power required by the compressor to achieve the desired pressure
ratio P2: P1 will be given by:
1
Power required by compressor C P
T1
P2
P1
m C
1 m
(4.28)
As already discussed earlier in this chapter, equation (A2.10) in Appendix 2 shows
is related to the fuel‐cell electrical power output, the average cell
that the parameter m
voltage and the air stoichiometry, i.e.,
3.58 10
m
7
Pe
Vc
(A2.10)
Substituting this relationship, electrical power Pe
air into equation (4.28) yields:
Compressor power 3.58 10
4
T1
m C
P2
P1
nIVc and the values of CP and γ for
0.286
1
In
(4.29)
Proton‐Exchange Membrane Fuel Cells
The effect of the loss due to the compressor can also be expressed as a voltage loss,
ΔVloss, simply by dividing the power given in equation (4.29) by the total current, I, and
for the number of cells in the stack, n, thus:
Vloss
3.58 10
4
T1
m C
P2
P1
0.286
1
(4.30)
The equations now provide a quantitative means of estimating whether a pressure
increase will improve the net performance of the fuel‐cell system. Equation (4.26)
provides the voltage gain by the fuel‐cell stack, and equation (4.30) can be used to
estimate the voltage loss due to the compressor.
It is possible to plot values of:
Net V
V gain
(4.31)
Vloss
for different values of P2/P1, and two examples are given in Figure 4.30, one case is designated
‘optimistic’, the other ‘realistic’. For these examples, the values of the various parameters
required in equations (4.26) and (4.30), i.e., C, T1, ηm,ηC and λ are given in Table 4.4.
0.02
‘Optimistic’ model
Net voltage change/V
0.015
0.01
0.005
0
1
2
3
4
5
6
7
Pressure ratio
–0.005
–0.01
–0.015
‘Realistic’ model
–0.02
Figure 4.30 Net voltage change that results from operating at higher pressure — for two different
PEMFC designs.
Table 4.4 Parameters for the examples given in Figure 4.30.
Optimistic model
Voltage gain constant (C), V
Inlet gas temperature, °C
0.10
15
Realistic model
0.06
15
Efficiency of drive for electric compressor (ηm)
0.95
0.90
Compressor efficiency (ηC)
0.75
0.70
Air stoichiometry (λ)
1.75
2.0
119
120
Fuel Cell Systems Explained
For the optimistic model, there is a net gain of about 17 mV per cell when the pressure
is boosted by a ratio of about 3, but the gain diminishes at higher pressures. For the
more ‘realistic’ model, however, there is always a net loss as a result of the higher
pressure. The power gained is always exceeded by the power needed to drive the
compressor. This shows clearly why operating at above atmospheric pressure is by no
means beneficial even with larger PEMFCs.
4.7.3
Other Factors
From the elementary analysis just given, one may wonder why pressurized operation
should be considered at all. The reason is that although it is the simplest to quantify, the
voltage boost is not the only benefit from operating at higher pressure. Similarly, the power
required by the compressor is not the only loss. High pressure can also enhance fuel
reforming. Whereas thermodynamics shows that hydrogen production by the steam
reforming of liquid hydrocarbons is favoured by operating at low pressures, the required
size and therefore the cost of the reactor hardware are reduced if the operating pressure is
increased. Humidification of the reactant air is also favoured by pressurization. Less water
is required to achieve the same level humidity in the cell at elevated pressures compared
with air at atmospheric pressure — refer back to Section 4.4.3 where equation (4.15) shows
that the humidity of the cathode exhaust air is dependent on the cell operating pressure.
With large fuel cells, the flow paths will be quite long and narrow. Therefore, the reactant
gases have to be pressurized to overcome frictional losses. A challenge for the fuel‐cell
system designer is in selecting a blower or compressor that matches the flow rate required
and the pressure drop imposed by the architecture of the stack. For example, other than
for simple PEMFCs of low power, an air blower or fan will always be required to overcome
the pressure drop through the cathode flow-fields of a stack. Such a fan has to be replaced
with a generally more expensive compressor for a system operating at pressure. From
a practical point of view, therefore, the extra size, weight and cost of high‐pressure
compressors compared with low‐pressure blowers have to be considered.
In the discussion thus far, it has been assumed that air supplies the cathode with
oxygen. There are, however, some fuel cells — notably, in space applications — that run
on pure oxygen from pressurized cylinders. In such systems, the operating pressure of
the stack will be chosen by balancing the advantage of the higher performance at
elevated pressure against the increased weight of the stack that is necessary mechanically
to withstand the high internal pressure. The optimum pressure will probably be much
higher than for air systems.
4.8
Fuel Types
4.8.1
Reformed Hydrocarbons
Up to now in this chapter, it has been generally assumed that the PEMFCs have been running
on pure hydrogen gas as the fuel and air as the oxidant. In small systems, this will usually
be the case. In larger systems, however, the hydrogen will frequently come from a fuel
processing or reforming system that produces carbon monoxide (CO) as a by‐product. A
prime example is the steam reforming reaction between methane and steam, i.e.,
CH 4 H2O
3H2 CO
(4.32)
Proton‐Exchange Membrane Fuel Cells
Whereas some of the high‐temperature fuel cells described in later chapters can use
this CO as a fuel, this does not apply to the PEMFC. Any CO in the fuel stream of a
PEMFC will be preferentially absorbed on the platinum catalyst in the anode electrode.
Consequently, hydrogen fuel is prevented from reaching the active platinum sites,
thereby inhibiting the oxidation reaction on the anode. Experience shows that a CO
concentration even as low as 10 ppm in the fuel gas degrades the performance of a
PEMFC. Therefore, if a reformed hydrocarbon is to serve as a fuel, the CO has to be
removed or at least reduced to a very low level. The extraction process is usually carried
out in several stages. Initially, CO and steam are passed over a catalyst that promotes
the water–gas shift reaction:
CO H2O
(4.33)
H2 CO2
Not all of the CO is converted by this reaction — an equilibrium point, governed by
the process conditions, is reached at 250°C, for example, the product gas from a shift
reactor will contain 1–2 vol.% of CO. Further process steps are therefore required for
reducing the concentration of CO to levels below a few ppm; these steps are described
in detail in Chapter 10. The shift reactor and additional processing steps add considerably
to the cost and size of a PEMFC system.
In some cases, the requirement to remove CO can be made somewhat less
demanding by the addition of small quantities of oxygen or air to the fuel stream that is
being fed to the PEMFC. At the catalyst sites on the fuel electrode, CO is converted
directly to CO2 by reaction with the oxygen. Reported results show, for example, that
adding 2 vol.% oxygen to a hydrogen gas stream containing 100 ppm CO eliminates the
poisoning effect. On the other hand, any oxygen not reacting with CO will certainly
react with hydrogen and thus waste fuel. Also, the method can only be used for CO
concentrations below about 100 ppm, which are not the levels found in the product
stream of a typical fuel reformer. In addition, the system required to feed precisely
controlled amounts of air or oxygen will be fairly complex, as the flow rate has to match
carefully the hydrogen supply rate.
Another important point to note is that the problem with CO intensifies with
hydrocarbons of increasing molecular length. The initial methane (CH4) reforming
reaction (4.32) produces three molecules of hydrogen. By contrast, the processing of a
fuel such as n‐octane (C8H18):
C 8H18 8H2O
(4.34)
17H2 8CO
results in a gas where the ratio of H2 to CO is now about 2 : 1.
4.8.2
Alcohols and Other Liquid Fuels
For any type of fuel cell, an ideal fuel would be a liquid that is already in regular use,
such as petrol or diesel. Unfortunately, these two fuels simply do not react at a sufficient
rate to warrant consideration for PEMFC systems. Possible alternatives to hydrogen in
a PEMFC are methanol and, to a lesser extent, ethanol; both are widely available
commercially. Methanol reacts at the anode of a PEMFC, albeit slowly, according to
the equation:
CH3OH H2O
6H
6e
CO2
(4.35)
121
122
Fuel Cell Systems Explained
Note that the methanol needs to be mixed with water and that six electrons are
produced for each methanol molecule and that the reaction does not directly produce
CO. This is the operational reaction of the DMFC which, together with other fuel cells
that operate directly on liquid fuels, is discussed further in Chapter 6.
The following section presents three typical applications of PEMFs. All the systems
employ stacks that operate with approximately ambient air pressure and use pure
hydrogen as the fuel and air as the oxidant.
4.9
Practical and Commercial Systems
4.9.1
Small‐Scale Systems
A class of PEMFC stacks with outputs of between a few watts and 1 kW is found in a
variety of applications, namely: (i) as battery chargers for portable electronic equipment
(e.g., mobile phones and laptop computers), (ii) for military use as personal power
sources and (iii) as stationary backup power supplies. Some of the smallest systems use
methanol and are described in Section 6.1, Chapter 6. Horizon Fuel Cells, in collaboration
with the associated company Horizon Energy Systems of Singapore, has championed
small, hydrogen‐fuelled PEMFC systems for several years and now markets a range of
air‐breathing stacks with outputs of 12 W to 1.0 kW for portable and educational
systems; examples are shown in Figure 4.31. The company also produces the ‘Mini‐pak’
(a)
(b)
(c)
Figure 4.31 Horizon Fuel Cell products: (a) 12‐W ‘H‐Series’, (b) 1‐kW ‘H‐Series’ and (c) ‘Mini‐Pak’ phone
charger. (Source: Reproduced with permission of Horizon Fuel Cells.)
Proton‐Exchange Membrane Fuel Cells
Figure 4.32 Mobile phone charger from Intelligent Energy (the ‘UPP’).
fuel cell for charging electronic devices; it uses an air‐breathing stack and a cartridge
that contains hydrogen stored as a hydride.
Following the introduction of the ‘Mini-pak’ into the US camping and outdoor markets
in the United States and Europe, Horizon Fuel Cells teamed up with Brunton to produce
the ‘Brunton Hydrogen Reactor’ for charging most pocket devices such as smartphones,
iPads, camera batteries, UV water purifiers, rechargeable lights and GPS units. A similar
product, known as the ‘UPP’, from Intelligent Energy in the UK is presented in Figure
4.32. Both products use small air-breathing fuel-cell stacks. The UPP device employs a
hydride cartridge (90.5 mm × 40 mm × 48 mm, weight 385 g) that can deliver 25 Wh of
energy. The stack is a 5‐W PEMFC that is able to produce up to 1000 mA at 5 V. One fuel
cartridge will therefore provide a smartphone with approximately five full charges, and it is
approved for carriage onboard aircraft. Each cartridge has a life of 9 years and is therefore
ready to meet any emergency well within the expected lifetime of the smartphone.
4.9.2
Medium‐Scale for Stationary Applications
Several companies are marketing PEMFC systems for backup or stationary power
systems, for example, for remote telecommunications towers and data centres. Systems
below about 5 kW such as those produced by PlugPower/Relion and Altergy employ
air‐cooled stacks. For the reasons given in Section 4.5.3, systems above 5 kW such as
those produced by Ballard/Dantherm, Hydrogenics and M‐Field are water cooled.
The following description of a Hydrogenics fuel‐cell power module (Figure 4.33) is
given as an illustration of a PEMFC product designed for stationary applications such
as for data centres. The module employs a water‐cooled stack of fairly conventional
design and is composed of 60 cells, each with an active area of 500 cm2, and bipolar
plates fabricated from a compression‐moulded carbon–polymer composite. The stack
produces about 12 kW at current of 350 A and a nominal voltage of 35–58 V. It is self‐
humidified, that is, there is no external humidification of either the fuel or air streams.
The essential balance-of-plant (BOP) items in the Hydrogenics power module is as
indicated in the schematic process flow diagram of Figure 4.28. Careful control of gas
flow rates and stack temperature (i.e., through flow of cooling water) keeps the stack at
123
124
Fuel Cell Systems Explained
Anode recycle compressor
Stack
Air filter and blower
Pressure control valve
Figure 4.33 Hydrogenics, rack‐mounted, ‘HYPM’ module, covers removed to show the stack and
balance‐of‐plant items. (Source: Reproduced with permission of Hydrogenics.)
optimum humidity. Recirculation of the anode exhaust gas helps to maintain even
humidity on the anode side of the stack. A differential pressure control valve ensures
that the pressure on the air side of the stack closely follows that on the fuel side.
There is a pump that circulates the hydrogen through the anodes of the stack, and
the fuel loop is ‘dead-ended’. A relief valve purges this line at intervals to prevent a
build‐up of contaminants within the anodes. Air is supplied via a blower, which is
regulated to provide the correct stoichiometry over the full operating regime of the
system. When the module is shut down, both the air and the fuel supplies are switched
off by solenoid valves. A small buffer vessel contains sufficient hydrogen so that any
remaining oxygen in the cathode air is self‐consumed by the system, and thereby
only inert gas remains in the shutdown state. This procedure is claimed to limit
degradation and prolong the life of the stack.
A process control system is coupled to the Hydrogenics power module and is embedded
with software both to monitor the performance of the stack and to adjust parameters,
such as hydrogen and air flow, in response to the electrical demand imposed. Other
process parameters that are fed to the controller include the stack temperature, cell
voltages and the pressure at the fuel side of the stack.
The Sankey diagram is a useful way to indicate the various energy flows and power
losses in a power‐generating system such as a fuel cell. The energy flows in an earlier
version of the Hydrogenics module are represented in the form of a Sankey diagram in
Figure 4.34. The diagram shows that only 10 kW of the 25.3 kW of energy embedded in
the hydrogen fed to the module appears as useful electrical power, i.e., the module has
an efficiency of 39% with respect to the lower heating value (LHV). Most of the energy
Proton‐Exchange Membrane Fuel Cells
10.9 kW heat removed
via cooling water
2.4 kW heat
lost via
exhaust gas 2 kW
waste heat
25.3 kW
hydrogen
12 kW DC power
from stack
10 kW total DC
electrical power
200 W
controller
700 W
1.06 kW UPS loss
DC-DC
converter
Figure 4.34 Sankey diagram of energy flows in a Hydrogenics fuel‐cell power module.
that is not converted to electricity is discharged as heat in the cooling water or exhaust
gas, or lost to the environment. Since the voltage produced by the stack varies according
to the load imposed, a DC–DC converter is employed to increase the voltage to a useful
and stable value. For stationary power applications, the DC output is usually converted
to AC so as to be compatible with the local network. The Sankey diagram in this case
shows that there are electrical losses associated with the DC–DC conversion and in
providing power to the system controller and battery uninterruptible power system
(UPS) that manages the module.
4.9.3 Transport System Applications
When Ballard Power Systems (BPS), a Canadian company, showcased its first PEMFC
stacks in the late 1980s, it became clear that this type of fuel cell was well suited for
application in electric vehicles. The high power density of PEMFC stacks, together with
zero‐emissions when fuelled by hydrogen, attracted companies such as DaimlerChrysler
and Shell who bought shares in BPS in 1994. New ventures were set up by Daimler to
develop the stacks and drivetrains for vehicles. Daimler built its first vehicle, the NECAR
(‘new electric car’), in 1994 and spent the next 20 years improving the fuel‐cell technology
through optimization of both the stack and the drivetrain components. The ensuing
developments led to the Mercedes B‐class F‐CELL which, in 2009, was the first fuel‐cell
car in series production.
Geoffrey Ballard, founder of the Canadian company, realized that buses provided a
unique opportunity to demonstrate his technology. Buses all refuel at a central depot,
125
126
Fuel Cell Systems Explained
they are amenable to be used with novel fuels, and they operate in cities where air
pollution is often a major issue. In August 1993, a 21‐seat Ballard bus carried its first
public passengers at the Commonwealth Games in Vancouver. This vehicle contained
only a small lead–acid starter battery as Ballard wanted to demonstrate that the fuel cell
could provide the motive power by itself. Seventeen years later, in 2010, BC Transit had
sufficient confidence in the technology to order twenty 12‐m, low‐floor, fuel‐cell buses
to carry participants between Vancouver and Whistler for the Winter Olympics. These
featured 130‐kW fuel‐cell stacks that were each supplied with hydrogen stored in a tank
at a pressure of 36 MPa. Hybridized with a nickel–metal-hydride battery, the buses had
a range of some 500 km.
Government‐supported demonstration programmes of fuel‐cell buses are now in
place throughout the developed world with manufacturers in Europe, Japan and North
America. Most recently, both China and India have become involved in such activity
through the development of their own PEMFC technology. Many of the fuel‐cell buses
on the road are hybrid vehicles and the latest model by Mercedes‐Benz, the Citaro
FuelCELL‐Hybrid, incorporates lithium batteries alongside the PEMFC stacks; one of
the buses is shown in Figure 4.35. As with previous Mercedes‐Benz buses, hydrogen for
the fuel‐cell stacks is stored at pressure in cylinders in the roof shell of the vehicle. The
number of cylinders required has been reduced from 9 to 7 (Figure 4.36) on account of
improved system efficiency and the use of the lithium‐ion batteries. These batteries
(which like the PEMFC stacks are water cooled) have a capacity of 27 kWh that is
sufficient to power the wheel‐hub electric motors at a constant 120 kW (165 hp). The
fuel consumption is 11–33 kg hydrogen per 100 km — i.e., 50% less in comparison with
its predecessor, the Citaro F‐CELL — and the range of the vehicle is 250 km.
Figure 4.35 Mercedes‐Benz Citaro FuelCELL‐Hybrid bus. (Source: Reproduced with permission of
Daimler.)
Proton‐Exchange Membrane Fuel Cells
Figure 4.36 Roof compartment of the Mercedes‐Benz Citaro FuelCELL‐Hybrid bus (showing seven
storage tanks and lithium batteries behind them). (Source: Reproduced with permission of Daimler.)
Fuel‐cell stacks for some of the early converted buses were situated where the engine
would be found in a diesel counterpart. The latest Citaro buses, however, have the fuel
cells located in the roof shell, at the rear of the bus behind the hydrogen cylinders. The
batteries sit between the hydrogen cylinders and the fuel‐cell stacks (see Figures 4.36
and 4.37). Thus all of the drivetrain is essentially mounted in the roof of the vehicle. Other
mechanical components required for operation of the bus, such as air‐conditioning
pumps, electric‐steering pump, air pump and inverter for auxiliaries, are placed in what
otherwise would be the rear engine compartment of a conventional diesel bus. This
placement allows easy access for servicing.
Vehicle manufacturers adopted Ballard stacks to demonstrate PEMFC technology in
cars. A 75‐kW design was the standard for most of the early Daimler fuel‐cell cars. As
confidence in the technology grew, however, automotive companies developed their
own stack technology. Examples include those built by General Motors, Honda (see
Figure 4.2), Hyundai, Nissan, PSA Citroen Peugeot, Toyota and Volkswagen. The Toyota
Mirai, launched in 2015, employs a 115‐kW stack that delivers power to a single
114‐kW electric motor. Hydrogen is contained at pressure in two tanks with a combined
volume of 122.4 L, and the manufacturer claims that this storage will give a range of up
to 650 km. The Hyundai ix35 fuel‐cell car is fitted with a 100‐kW stack and promises a
range of 594 km from one charge of hydrogen at 70 MPa.
In the United Kingdom, Intelligent Energy is developing fuel cells with various
companies, which include the motorbike manufacturer Suzuki, and has announced their
own innovative 100‐kW water‐cooled system for vehicles; see Figure 4.38. As explained by the
company, the 100‐kW platform takes full advantage of Intelligent Energy’s stack technology
127
128
Fuel Cell Systems Explained
Figure 4.37 Roof compartment of the Mercedes‐Benz Citaro FuelCELL‐hybrid bus (showing the fuel
cell stacks and associated BOP). (Source: Reproduced with permission of Intelligent Energy.)
(a)
(b)
Figure 4.38 Intelligent Energy 100‐kW fuel‐cell system for vehicles: (a) water‐cooled stack and
(b) packaged system. (Source: Reproduced with permission of Intelligent Energy.)
that offers a power density of 3.5 kW L−1 and a specific power of 3.0 kW kg−1, while being
engineered for low‐cost, high‐volume series production. The key to this performance is said
to be the proprietary, evaporatively cooled (EC) technology. The stack employs metal
separator plates, and compared with conventional liquid‐cooled fuel‐cell stacks, the EC
design is said to remove the need for individual cooling channels between each cell and
thereby delivers considerable advantages in terms of the reduction in both stack volume and
mass. The technology also indicates that there continues to be room for innovation in PEMFC
technology, which bodes well for the future of applications such as fuel‐cell vehicles.
Proton‐Exchange Membrane Fuel Cells
4.10
System Design, Stack Lifetime and Related Issues
Many years of research have shown that durable, high‐performance and low‐cost
PEMFCs can be achieved through the appropriate combination of materials, design and
operating conditions. Investigations have also helped to identify the means by which
the following modes of cell degradation may eventuate.
4.10.1
Membrane Degradation
Mechanical degradation may be caused by swelling of the membrane through, for
example, poor water management. The membrane can also breakdown as a result of
chemical reaction by foreign elements such as precipitated platinum from the
catalysts or iron from metal bipolar plates. Degradation can occur through peroxide
formation by the ORR. It has been shown that the aggressive action of peroxide is
accelerated by the presence of iron and is believed to be due to the generation of
hydroxyl (OH–) and hydroperoxyl (HOO–) radicals, which attack the acidic moieties
in the polymer membrane.
In summary, to avoid degradation of the membrane, it is important to address at least
one of the following actions:
●
●
●
●
●
●
Reduce peroxide generation or accelerate in situ peroxide decomposition.
Remove or passivate iron and other undesirable metal contaminants.
Enhance the oxidative stability of the membrane.
Improve water management.
Reduce time spent at >0.9 V.
Ensure membrane is sufficiently hydrated, e.g., by limiting high‐temperature operation.
4.10.2
Catalyst Degradation
On the cathode side, sintering or dissolution of platinum may be reduced by alloying
with other elements, e.g., cobalt and iridium. More stable catalyst supports, such as
graphitized carbon, are also required. Operational benefits include introducing and
maintaining hydrogen at the anode when the cell is ‘off ’, i.e., no load, and shorting or
applying an immediate load on cell start‐up to remove air from the cathode. Care also
needs to be taken to extract any potential catalyst poisons from the fuel and air streams.
4.10.3
System Control
Much effort has been devoted to designing appropriate control technology to ensure
that a fuel‐cell system operates under conditions that best prolong the lifetime of the
MEA. This task requires the performance of the stack to be accurately monitored, and
the most convenient method is to measure the voltages of individual cells, or groups of
cells, in real time. A microprocessor or a programmable logic controller (plc) system can
be used to read, record and analyse the voltages in response to changes in demand by the
load. The controller can actuate valves and other devices to change operating parameters
such as gas flows, humidifier temperature, and system pressures. Consequently, the
controller may, for example, maintain the correct stoichiometries of air or oxygen to
enable the cell voltages to remain constant within narrowly defined limits. If the voltage
129
130
Fuel Cell Systems Explained
of one cell falls significantly, an alarm situation can be enunciated to indicate that some
remedial action may be required to prevent the one cell from causing a reduction in the
voltage of the whole stack and thereby accelerate stack degradation. Similar provisions
are embodied in the battery management systems of advanced lithium‐ion batteries.
If the PEMFC system employs several stacks in modules, the plc will oversee the
operation of the whole system to ensure that each module is operating cohesively. For
example, when the fuel‐cell system receives a start‐up signal from an operator, the
controller gives instructions so that the system follows a pre‐established boot‐up
procedure that starts each module at an appropriate time. The strength of using a plc in
this role is that, by programming suitable algorithms, it can be employed to detect stack
malfunctions, such as unusual variations in cell voltage due to fluctuating rates of gas
flow caused by blocked channels in the bipolar plates. To enable stacks to be integrated
within vehicle systems, the PEMFC controller is usually made to communicate with
other components via a controller area network (CAN bus).28
4.11
Unitized Regenerative Fuel Cells
A unitized regenerative fuel cell (URFC) is a reversible cell that is able to operate as a
conventional fuel cell and, in regenerative mode, as an electrolyser. When in electrolyser
mode, the URFC generates hydrogen and oxygen by the electrolysis of water (see
Section 10.8, Chapter 10). Both modes are carried out with the same fuel‐cell stack. In
comparison with a separate fuel cell and electrolyser, the combination of these duties
in the same hardware holds several advantages, such as lower capital cost, simpler
structure, higher specific energy and no need for auxiliary heating. Although both the
AFC and SOFC have received some attention as reversible fuel cells, systems based on
PEMFC stacks are the most mature. Designs of URFC have already been employed in
aerospace applications.
Rechargeable secondary batteries are widely used for energy storage purposes due to
their high round‐trip efficiency (around 80%), but they suffer from some obvious
drawbacks. The durability of lead–acid batteries is not very satisfactory when faced
with deep cycling, and their specific energy is constrained by the heavy weight. Lithium‐
ion batteries promise to be much more durable with respect to cycling, but are subject
safety issues. Redox flow batteries (RFBs), as described in Section 1.7.2, Chapter 1, have
attracted interest because they provide the means of decoupling energy storage capacity
and rated power. By enlarging the electrolyte storage tank, the capacity can easily be
increased, while the rated power can be enhanced by using electrodes of greater area or
through stacking. On the other hand, due to the bulk electrolyte solution contained in
the system, the specific energy of RFBs is generally much lower.
28 A controller area network (CAN) bus is a vehicle electronic serial bus standard that is designed to allow
microcontrollers and other electrical or electronic devices to communicate with each other in applications
without the need for a host computer. CAN bus is a message‐based protocol and, although designed
originally for automotive applications, it is used in many other contexts. The modern automobile has many
electronic control units for various subsystems. Typically, the biggest processor is the engine control unit.
Others are used for items such as transmission, airbags, anti‐lock braking/anti‐skid braking system (ABS),
cruise control, electric power steering, audio systems, power windows, doors, mirror adjustment and
battery recharging systems for hybrid/electric cars.
Proton‐Exchange Membrane Fuel Cells
Box 4.1 Pure oxygen versus air in a PEMFC
Running a PEMFC with oxygen rather than air as the cathode gas markedly improves cell
performance by virtue of the following three effects:
1) The ‘no loss’ open‐circuit voltage rises on account of the increase in oxygen partial
pressure, as predicted by the Nernst equation; see Section 2.5, Chapter 2.
2) The activation overpotential reduces through better use of catalyst sites; see
Section 3.4.3, Chapter 3.
3) The limiting current increases and thus reduces the mass-transport or concentration
overpotential losses. This benefit is due to the removal of the nitrogen gas, which is a
major contributor to such losses at high current densities; see Section 3.7, Chapter 3.
Depending on the design of PEMFC, a change from air to oxygen can increase the power
of the stack by about 30%. In particular, a stack with poor reactant air flow will benefit
more from a switch to oxygen. For a URFC system that involves the storage of oxygen and
hydrogen, the use of pure oxygen has a significant impact; it may increase the round‐trip
efficiency from a typical 35 to 50%.
As with flow batteries, URFCs also store the fuel and oxidant, generally H2 and O2,
externally in separated gas tanks and therefore offer the ability to decouple storage
capacity and output power. By contrast, however, their specific energy is much higher
than that of RFBs, i.e., about 0.4–1.0 kWh kg−1 (including the mass of the hydrogen and
oxygen gas tanks29) compared with 0.01–0.02 kWh kg−1 for a vanadium redox battery.
In addition, URFCs can be totally charged and discharged without damaging the
durability of the fuel cell. These advantages have made URFCs very competitive against
secondary batteries and flow batteries. On the debit side, however, URFCs generally
achieve lower round‐trip efficiency than batteries (typically below 40%) due to the
sluggish reactions for oxygen evolution and oxygen reduction. Efficiency can be
increased if hydrogen and oxygen are stored and used (see Box 4.1). The low efficiency
would also be more tolerable if the URFC could be employed in a cogeneration system,
where the heat that is generated in fuel‐cell mode could be harnessed. Other issues
such as high cost, hydrogen storage, and relatively low technology readiness, have also
hindered their exploitation.
In practice, there are other technical issues concerning PEMFC‐based URFCs. These
mainly concern the bifunctional catalyst that has to service both the ORR and the
oxygen evolution reaction (OER). To date, most of the bifunctional catalysts utilized in
URFCs are based on noble metals. Platinum (Pt), the preferred catalyst for the ORR, is
not suitable for the OER. Moreover, the preferred catalysts for the OER, such as
ruthenium (Ru), iridium (Ir) and the oxides of the two metals, are not suitable for the
ORR. Consequently, a compromise has to be made from the combination of these candidate metals and oxides that delivers the best performance, as composite catalysts. The
combination of Pt and Ir or its oxides is currently the preferred choice for a bifunctional
29 Mitlitsky F, Myers B, and Weisberg AH, 1988, Regenerative fuel cell systems, Energy Fuels, vol. 12,
pp. 56–71.
131
132
Fuel Cell Systems Explained
catalyst. Numerous studies on the optimization of the two metals (e.g., elemental ratio,
method of catalyst preparation, microstructure) have been conducted.
Carbon, which is the preferred catalyst support material for PEMFCs, is less suitable
for URFCs because, on the oxygen side of the cell, carbon corrosion is promoted under
electrolysis conditions. For this reason other support materials such as titania, titanium
carbide or nitride have been investigated. As noted in Section 4.3.2, the study of catalyst
materials, particularly non‐precious metals, for the cathode in PEMFCs is a very active
research area, one from which URFCs could also benefit. Similarly, the carbon‐based
GDL that is employed in the PEMFC is not suitable for the URFC, and alternatives are
under investigation.
Despite these issues, several developers have produced URFCs that employ PEMFC‐
type stacks and include the following:
●
●
●
●
Distributed Energy Systems (Connecticut, USA) has constructed a multi‐kW, closed‐
loop, lightweight URFC for high‐altitude airships, that can generate pressurized
hydrogen and oxygen electrochemically without mechanical compression.30
The NASA Glenn Research Center demonstrated a closed‐loop URF for a solar electric aircraft in 2006.31 The system could store the input electrical energy and output a
steady electrical power of 5 kW for at least 8 h.
IHI (Japan) collaborated with Boeing to develop a URFC for aircraft auxiliary power
units (APUs).
Demonstration systems have been produced in the United States by Giner Inc.,
Lynntech, Lawrence Livermore National Laboratory and Proton Energy Systems Inc.
Further challenges for the URFC arise from the management of water consumption in
electrolysis mode or production in fuel‐cell mode. Water management is therefore even
more complex than that for the PEMFC as described in Section 4.4.
Further Reading
Barbir, F, 2012, PEM Fuel Cells: Theory and Practice, Academic Press, Waltham, MA.
Behling, N, 2012, History of proton exchange membrane fuel cells and direct methanol fuel
cells, in Fuel Cells: Current Technology Challenges and Future Research Needs, pp.
423–600, Elsevier, Amsterdam.
Koppel, T, 1999, Powering the Future – The Ballard Fuel Cell and the Race to Change the
World, John Wiley & Sons, Inc., New York.
Gasteiger, HA, Baker, DR, Carter, RN, Gu, W, Liu, Y, Wagner FT and Yu PT, 2010,
Electrocatalysis and catalyst degradation challenges in proton exchange membrane fuel
cells, in Stolten D (ed.), Hydrogen and Fuel Cells, Fundamentals, Technologies and
Applications, pp. 3–16, Wiley‐VCH, Weinheim.
Reijers, R and Haije, W, 2008, Literature review on high temperature proton conducting
materials: Electrolyte for fuel cell or mixed conducting membrane for H2 separation,
30 Funding, demo for regenerative fuel cell, 2004 Fuel Cells Bulletin, 2004, pp. 7–8.
31 Bents, DJ, Scullin, VJ, Chang, BJ, Johnson, DW, Garcia, CP and Jakupca, IJ, 2006, PEM hydrogen‐oxygen
regenerative fuel cell development at NASA Glenn Research Center, Fuel Cells Bulletin, vol. 2006, pp. 12–14.
Proton‐Exchange Membrane Fuel Cells
Report no. ECN‐E‐‐08‐091, prepared under the KIMEX project no. 7.0330, ECN
Research Centre, Petten, the Netherlands.
Zhang, J, Xie, Z, Zhang, J, Tang, Y, Song, C, Navessin, T, Shi, Z, Song, D, Wang, H,
Wilkinson, DP and Liu, ZS, 2006, High temperature PEM fuel cells, Journal of Power
Sources, vol. 160(2), pp. 872–891.
Wang, Y, Leung, DYC, Xuan, J and Wang, H, 2016, A review on unitized regenerative fuel
cell technologies, part‐A: Unitized regenerative proton exchange membrane fuel cells,
Renewable and Sustainable Energy Reviews, vol. 65, pp. 961–977.
133
135
5
Alkaline Fuel Cells
5.1
Principles of Operation
The basic chemistry of the alkaline fuel cell (AFC) has been explained in Figure 1.4,
Chapter 1. The reaction at the anode is:
2H 2
4OH
4 H2 O 4 e
E
0.282 V
(5.1)
where E° is the standard electrode potential. The electrons released travel round the
external circuit to the cathode, where they react to form new OH ions, i.e.,
O2
4e
2 H2 O
4 OH
E
0.40 V
(5.2)
References to the AFC can be traced back at least to 1902,1 but it was the work of F. T.
(Tom) Bacon, first at the University of Cambridge (1946–1955) and then at Marshall of
Cambridge Limited (1956–1961), which led to the first practical demonstration of the
technology. The Bacon cell was adopted for the Apollo space programme — an example
is shown in Figure 5.1 — and this created the general impression that the AFC was an
expensive and specialized system. Later, however, Kordesch at Union Carbide
(Cleveland, Ohio) and Justi and Winsel at Siemens (Erlangen, Germany) showed that an
atmospheric pressure hydrogen–air AFC could work very effectively, with the proviso
that carbon dioxide (CO2) must not be present in the fuel or oxidant unless means of
either purification or replacement of the electrolyte solution were included. Experimental
AFCs were tested for their ability to power agricultural tractors, cars, offshore navigational
equipment, boats, forklift trucks and various other applications during the 1960s and
early 1970s. Although many of the systems worked reasonably well as demonstrations
of proof‐of‐concept, issues such as cost, reliability, ease of use, ruggedness and safety
proved to be a challenge. During the 1980s and 1990s, prospects appeared poor for the
AFC when compared with other emerging fuel cells. Consequently, research was scaled
down so that by the close of the century only a couple of companies were actively
working on AFCs. According to most analysts, the emergence of the proton‐exchange
membrane fuel cell (PEMFC) heralded the final demise of the AFC, especially when a
decision was taken in 1997 to replace the system that had been used for the Space
Shuttle Orbiter vehicles with PEMFCs for future missions.
1 Reid, JH, 1902, US Patent no. 736 016 017.
Fuel Cell Systems Explained, Third Edition. Andrew L. Dicks and David A. J. Rand.
© 2018 John Wiley & Sons Ltd. Published 2018 by John Wiley & Sons Ltd.
136
Fuel Cell Systems Explained
Figure 5.1 Alkaline fuel‐cell system (1.5 kW)
employing 32 circular fuel cells (200 mm diameter) as
used in the Apollo spacecraft. The cells were in the
lower container that was purged with nitrogen to
remove waste heat. The stack area of 465 cm2 gave
0.86 V at 470 mA cm−2 (4.0 kW m−2). Three of these
units were connected in parallel to provide
redundancy, each weighed 109 kg. The fuel cells
provided the electrical power, as well as much of the
potable water, for the craft that took mankind to the
moon. (Courtesy of International Fuel Cells.)
Despite the lack of interest in the AFC by many developers, it should be pointed out
that this technology does offer some technical advantages over the more successful
PEMFC and phosphoric acid fuel cell (PAFC) alternatives. The activation overpotential
at the cathode of an AFC is generally lower than that in acid fuel cells, and the electrode
reactions are faster. It is therefore not essential to use platinum‐based catalysts in the
AFC. Furthermore, the efficiency of electrical generation by the AFC is generally greater
than that of the PEMFC on account of the lower overpotential at the cathode. Indeed it
was the high efficiency of energy conversion — typically, 70% (LHV) — that led to the
National Aeronautics and Space Administration (NASA) selecting the AFC for deployment
in the US space programme.
Conventional AFCs use an alkali electrolyte dissolved in water. Sodium hydroxide
(NaOH) and potassium hydroxide (KOH) — the most abundant and cheapest alkaline
hydroxides — were the prime candidates for the early AFCs. Unfortunately, however,
CO2 present either in the fuel or oxidant streams can react with such hydroxides and
cause the formation of potassium or sodium carbonate in the electrolyte solution, for
example:
2KOH CO2
K 2 CO3
H2 O
(5.3)
Alkaline Fuel Cells
Such a reaction has the following adverse impacts:
●
●
●
●
●
Decrease in the OH− concentration in the electrolyte solution, thereby interfering
with the kinetics of the cell reaction.
Increase in the viscosity of the electrolyte solution, thereby resulting in a slower
diffusion rate and lower limiting currents.
Precipitation of carbonate salts in the porous electrode, thereby reducing mass transport.
Reduction in oxygen solubility.
Decrease in the conductivity of the electrolyte solution.
The end result is a severe deterioration in cell performance. Of the two candidate
hydroxides, KOH is generally preferred because its carbonate is much more soluble in
water than the sodium counterpart. Degradation of the AFC by CO2 was a significant
contributor to the abovementioned declining interest in AFCs. Nowadays, however,
there is a renewed enthusiasm for the AFC that has been brought about by an improved
understanding of the influence of CO2 on cell performance and by the emergence of
anionic polymer membranes to replace the traditional electrolyte solutions.
5.2
System Designs
5.2.1
Circulating Electrolyte Solution
The design of AFC in which the electrolyte solution is circulated was pioneered by Bacon
and subsequently used by Pratt and Whitney (later International Fuel Cells) in the 1950s for
the Apollo missions. A schematic representation of essentials of the system is presented in
Figure 5.2. In the Bacon design, an aqueous solution of the electrolyte (typically 33 wt.%
KOH) is pumped through the fuel cell. Hydrogen is supplied to the anode but must be
circulated as water is produced at this electrode.2 When pressurized to 500 kPa, the
cell operates at 200°C, and therefore the product water is actually steam and has to be
condensed out from the circulating hydrogen. For the Apollo missions, the pressure was
reduced to 330 kPa, and the concentration of the electrolyte solution was increased to
85 wt.% KOH. As shown in Figure 5.2, hydrogen is supplied from a compressed gas cylinder.
Unlike the fuel‐cell systems employed in the US space programme that were supplied
with pure oxygen, stationary AFC power plant invariably operates with air as the oxidant.
This can be supplied with a blower since the pressure drop over a cell is usually low
(around 2.0 kPa). To avoid cell degradation, a scrubber is installed in the air line to remove
CO2 to a level well below 50 ppm. For small‐scale AFCs, the scrubber can simply be a
vessel that contains soda lime, which has to be discarded once it has absorbed its full
capacity of CO2. More substantial systems can use a regenerative scrubber that employs an
amine‐based material in two parallel reactors that can be alternated between absorption
and desorption cycles. One reactor absorbs the CO2 present in the air flowing towards
the stack. Meanwhile, in the other reactor, the exhausted amine material, which has been
used previously for treatment of the incoming air, is rejuvenated by desorbing the CO2
with the excess air that is exiting the stack. The CO2 is weakly bound to the amine material and can be released by simply raising the temperature of the reactor.
2 As with PEMFCs, it is possible to run the AFC dead ended, but this is not preferred as the anode has to be
purged frequently to remove the product water and any contaminants.
137
138
Fuel Cell Systems Explained
Air pump
Air in
Ejector circulator
Carbon dioxide
removal
–
H2
H2
A
n
o
d
e
H2
E
l
e
c
t
r
o
l
y
t
e
+
Electrical
power
output
C
a A
t i
h r
o
d
e
Cooling air
Hydrogen cooler and
water condenser
Coolant circulation pump
Air out
Electrolyte
cooler
Electrolyte
circulation
pump
Figure 5.2 Diagram of an alkaline fuel cell with circulating electrolyte solution, which also serves as a
coolant for the fuel cell.
The disadvantages of a circulating electrolyte solution lie in the extra equipment that
is required in the form of pumps and a cooler. The pipework necessary to achieve the
circulation is prone to leak on account of the low surface tension of the aqueous KOH.
It is also a challenge to design a system that will function in any orientation. In a facility
that employs multicell stacks, the design must be such that the circulating electrolyte
solution does not provide an unwanted current path between cells.3
Systems with circulating electrolyte solutions do have advantages, however, and the
principle benefits are as follows:
●
●
The circulating electrolyte solution can serve as a cooling system for the fuel cell.
The electrolyte solution is continuously stirred and mixed. Reactions (5.1) and (5.2)
show that twice as much water is produced at the anode as is consumed at the cathode.
Without intervention, this will result in the electrolyte solution becoming too
3 This unwanted ‘internal’ or ‘shunt’ current can be determined by measuring the hydrogen consumption at
open circuit. For example, in the cells used by Kordesch (1971), the internal current density was found to be
about 1.5 mA cm−2.
Alkaline Fuel Cells
●
●
●
●
concentrated at the cathode — in fact, so concentrated that it solidifies. Stirring
reduces this problem.
Excess electrolyte solution can be stored in a vessel external to the stack. This
solution can be heated, if necessary, to drive off any additional water that has been
absorbed.
It is comparatively straightforward to pump out all the electrolyte solution and replace
it with a fresh solution.
Start‐up and shutdown are both simple — for a cold start‐up, only the reservoir holding
the electrolyte solution needs to be heated, rather than the whole stack.
The cell can be monopolar, which enables a stack design that is easier to build than
one that employs bipolar plates to interconnect the cells. Moreover, there is greater
flexibility to configure a stack in terms of the desired voltage and current.
Although monopolar designs may be easier to build than bipolar versions, edge
collection of current can lead to lower performance as a result of the cumulative resistance, or voltage drop, between the centre and edges of an electrode. The consequence
is a lower average current density over the whole electrode surface. This adverse effect
becomes more serious as stacks are scaled up.
Following the successful application of the AFC in spacecraft, a few companies have
pursued the technology for other applications; almost universally the cells have operated
with circulating electrolyte solutions. In the United States, Allis‐Chalmers had worked
on the technology during the 1960s and Union Carbide continued some work in 1970s.
Fuji Electric was the only Japanese company to support AFC technology for an appreciable
period. Consequently, it was left mainly to European companies — notably, Siemens
and later Elenco — to address the challenges of the AFC in the final years of the century.
Elenco located in Belgium was owned by the Belgium company Bekaert and the Dutch
State Mines until 1995, when its financing ended. Together with partners in the
Netherlands and France, Elenco built AFC systems for articulated city buses under a
EUREKA project that was supported by the European Union from 1991 to 1994. The
work showcased 40‐kW AFC systems. The rise of interest in the PEMFC in the 1990s,
however, signalled the end of Elenco and in a last‐bid effort to rescue the AFC, the
company was taken over by Zetek, a UK technology venture. Further interest was
generated by Zetek who retrofitted several vehicles, which included a London taxi, with
AFC systems that were based on the Zetek Mk2 stack illustrated in Figure 5.3a. Each
stack consisted of a series–parallel configuration of 24 individual cells that delivered a
power output of 434 W and an output current of 108 A at 4 V. Current densities of up to
120 mA cm−2 were measured at an operating temperature of 70°C. The stack was
normally run at 100 mA cm−2 at an average cell voltage of 0.67 V.
Rebranded as Zevco, Zetek continued its AFC development until 2001. The Elenco–
Zetek–Zevco technology worked extremely well and showed multiple year lifetimes.
Unfortunately, the company was caught in the turmoil after the September 11 attacks
and the expected financial closure with investors failed to materialize due to the
disruption of trading on international markets. In the fall‐out following the collapse of
Zetek, new companies were established to take on the intellectual property and move
the technology forward. These include AFC Energy (originally Eneco Ltd.) and, more
recently, Cygnus Atratus. The AFC Energy systems have a bipolar stack arrangement
that has been adapted from the original Zetek monopolar stack design. Low‐cost polymer
139
140
Fuel Cell Systems Explained
(a)
(b)
Fuel-cell stack
Figure 5.3 Alkaline fuel cells: (a) Schematic of Zetek Mk2 stack and (b) Intensys–Vito 6‐kW system.
(Source: Reproduced with permission of Cygnus Atratus.)
frames are used in both the AFC Energy and Cygnus Atratus technologies. In 2008, the
Flemish Technical Institute, VITO, in collaboration with Intensys introduced 6‐kW
systems that are also based on the Elenco–Zetek technology; an example is shown in
Figure 5.3b.
5.2.2
Static Electrolyte Solution
An alternative design of AFC, in which each cell in the stack has its own, separate,
electrolyte solution that is held in a matrix material between the two porous gas‐diffusion
electrodes, is shown schematically (Figure 5.4). The arrangement is clearly less complex
than that required for a circulating electrolyte solution and, as will be seen in Chapters
7 and 8, is similar to that used in the PAFC or the molten carbonate fuel cell (MCFC).
Furthermore, with a static electrolyte solution, the AFC stack can be used in any
orientation, and there is no risk of internal short circuits arising, as is the case with a
circulating electrolyte solution.
It was these key advantages of simplicity of design and the ability to work in any
orientation that led to the static electrolyte solution AFC that was produced by United
Technologies Corporation for the Space Shuttle Orbiter (Figure 5.5). The state‐of‐the‐art
alkaline fuel‐cell stacks in the Orbiter were rectangular (38 × 114 × 35 cm), weighed
118 kg and produced a peak power of 12 kW at a minimum of 27.5 V (end of life) and an
average power of 7 kW. The stacks operated at a similar pressure to the Apollo versions
(400 kPa) but at a lower temperature (85–95 vs. 200°C). Unfortunately, the lower
operating temperature necessitated the use of Pt catalysts.4
There are, however, some challenges for this type of AFC with respect to the durability
and the robustness required for commercial terrestrial applications. As the electrolyte
solution matrix can neither be removed nor be completely replaced once a cell has been
4 For the Orbiter fuel cell, gold‐plated nickel electrodes were employed onto which the catalyst was deposited.
The catalyst loading on each electrode was 20 mg cm−2 of Au–Pt alloy on the cathode and 10 mg cm−2 Pt on
the anode.
Alkaline Fuel Cells
Electrical
power
output
–
+
Ejector circulator
O2
H2
H2
A
n
o
d
e
H2
Cooling air
E
l
e
c
t
r
o
l
y
t
e
O2
C
a
t
h
o
d
e
Next
cell
O2
Hydrogen cooler and
water condenser
Coolant
also
flows
through
stack
Coolant circulation pump
Figure 5.4 Alkaline fuel cell with static electrolyte solution held in a matrix. The system uses pure
hydrogen and oxygen, e.g., as employed in spacecraft.
Figure 5.5 Alkaline fuel‐cell module used in Space Shuttle Orbiter.
141
142
Fuel Cell Systems Explained
assembled, any impurities or carbonates formed within the electrolyte solution will
inevitably accumulate. This can drastically reduce the cell performance. The electrolyte
solution also cannot be used for cell cooling although this may be achieved via the phase
change of water to steam, through the evaporation of water in the anode and/or cathode
gas streams. Alternatively a separate cooling system may be employed, as illustrated in
Figure 5.4, which was the approach taken for the Apollo and Orbiter spacecrafts.
The Apollo AFCs were cooled using a mixture of ethylene glycol and water, as is used in
car engines. In the Orbiter systems, the cooling fluid was a fluorinated hydrocarbon
dielectric liquid.
The system represented by Figure 5.4 uses pure oxygen at the cathode, though this is
not obligatory for a matrix‐held electrolyte solution. As in the design that operates with
a pumped electrolyte solution (see Figure 5.2), the hydrogen is circulated to remove the
product water. In spacecraft systems, the product water is used for drinking, cooking
and cabin humidification. Water management is, however, an issue and essentially is
similar to that for PEMFCs, though ‘inverted’ in that water is produced at the anode and
removed from the cathode. (In the PEMFC, water is produced at the cathode and
removed from the anode by electro‐osmotic drag, as explained in Section 4.4, Chapter 4.)
The AFC system must be designed so that the water content of the cathode region is
kept sufficiently high by diffusion from the anode. In general, the problem of water
management is much less severe than with the PEMFC. For a start, the rise in the
saturated vapour pressure of KOH solution with temperature is less rapid than that
shown by pure water, as will be discussed in Section 5.4. Accordingly, the rate of
evaporation is much slower.
In the earliest forms of AFCs with static electrolyte solution, the KOH solution was
held in a matrix made of asbestos that had excellent porosity, strength and corrosion
resistance. Given recognition of the health hazards associated with the deployment of
asbestos, alternative materials were developed for spacecraft. For instance, butyl‐
bonded microporous potassium titanate [(K2O)x•(TiO2)z z/x ≈ 8] or K2TinO(2n+1)
(n = 4.0–11.0) was used in the space shuttle fuel cell. Ceria and zirconium phosphate
have also been proposed, but as yet it appears that no substitute has proved to be
universally acceptable for the porous matrix. In addition, for terrestrial service, renewal
of the electrolyte solution from the matrix must be possible given that the problem of
CO2 contamination is bound to occur.
For other applications, the use of AFCs with static electrolyte solution may be
overtaken by cells that employ anion‐exchange membranes, as described in Section 5.2.4.
5.2.3
Dissolved Fuel
A fuel cell that operates with dissolved fuel is unlikely to be employed for serious
power generation but is included here as the design is the simplest to manufacture.
The dissolved fuel AFC, in particular, was popular for demonstrating the operating
principle of fuel cells and featured in early textbooks, before the widespread availability
of small‐scale educational PEMFC systems for schools and colleges. The underlying
concept is in Figure 5.6. The KOH electrolyte solution is mixed with a fuel, such as
hydrazine, ammonia or sodium borohydride. The fuel anode is along the lines
discussed in Section 5.3.4, with a platinum catalyst. The fuel is also fully in contact
with the cathode. Whereas this would markedly increase the severity of the ‘fuel
Alkaline Fuel Cells
Electrical
power
output
–
+
Waste
gasses
Air cathode
Electrolyte
and fuel
mixture
Fuel anode
Figure 5.6 Schematic representation of a dissolved fuel AFC, arguably the simplest of all types, it has
a selective catalyst on the cathode that does not react with the fuel. An alternative design has a
membrane within the electrolyte solution that isolates the fuel from the air cathode, but adds to cost
and complexity.
crossover’ problem (discussed in Section 3.5, Chapter 3), it is of no consequence here
as the cathode catalyst is not platinum and therefore the rate of reaction of the fuel is
very low. Furthermore, there is only one seal that could leak, namely, a very low
pressure joint around the cathode. The cell is re‐fuelled simply by adding more fuel to
the electrolyte solution.
Hydrazine, H2NNH2, is an ideal fuel for this type of cell because it dissociates into
hydrogen and nitrogen at the anode; the hydrogen that is formed reacts according to
equation (5.1).
Sodium borohydride (NaBH4) can also be used as a fuel. In Chapter 11, this compound
will be considered as a material for hydrogen storage. As a fuel, it can be dissolved in the
AFC electrolyte solution, and it reacts at the anode according to:
NaBH 4
8OH
NaBO2
(5.4)
6H2 O 8e
The impressive fact to note is that eight electrons are formed by this reaction for just
one molecule of fuel. Even more interesting is the large change in Gibbs free energy
(expressed as ∆ g f kJ mol−1, see Section 2.1, Chapter 2) and therefore the high reversible
voltage (Vr) of the cell. The reaction of air at the cathode reaction is exactly the same as
for the hydrogen fuel cell, i.e., equation (5.2). The overall reaction is thus:
(5.5)
NaBH 4 2O2 NaBO2 2H2 O
For this reaction:
920.7
Gf
2 237.2
123.9
1271.2 kJ mol
1
(5.6)
Therefore, from equation (2.9), Chapter 2:
Vr
Vr
g
f
(2.9)
zF
gf
zF
1271.2 103
8 96 485
1.64 V
(5.7)
143
144
Fuel Cell Systems Explained
This theoretical voltage is significantly higher than that obtained with hydrogen, and
at eight electrons per molecule, it indicates a fuel of remarkable potency. Unfortunately,
the voltage actually obtained with a borohydride fuel cell is not so different from
that with a hydrogen cell, because the catalysts that facilitate the direct borohydride
oxidation, reaction (5.4), also promote the following hydrolysis reaction:
NaBH 4
2H 2 O
NaBO2
4 H2
(5.8)
This reaction was the main reason for the abandonment of the technology in the
1960s. The electrodes available at the time were not able to utilize the hydrogen
effectively, so the loss of hydrogen via reaction (5.8) made the borohydride cell
inefficient. This is not the case with modern electrodes, which, even with low platinum
loadings, will promote direct hydrogen oxidation. Furthermore, if the concentration of
borohydride in the electrolyte solution is low, the rate of reaction (5.8) is reduced
significantly with the net effect of an improvement in cell voltage. As a fuel, sodium
borohydride is an expensive but convenient means of providing hydrogen. Further
discussion of borohydride fuel cells is given in Section 6.5, Chapter 6.
5.2.4
Anion‐Exchange Membrane Fuel Cells
In contrast to the PEMFC, the AFC exhibits facile kinetics for both the anode and
cathode reactions. Consequently, cheaper non‐noble metal catalysts can be used in the
electrodes. As noted in the previous sections, however, the AFC has a significant
drawback in that degradation of both the electrolyte solution and the electrodes can
occur through the formation of carbonate/bicarbonate (CO32−/HCO3−) via reaction
between OH− ions and CO2 in the oxidant gas stream.
A more recent variant of the AFC is the anion‐exchange membrane fuel cell (AMFC),
in which the KOH electrolyte solution is replaced by a solid alkaline‐electrolyte
membrane (AEM).5 The AEM is a polymer material, and the fuel cell is effectively
an alkaline analogue of the PEMFC. The AMFC thus retains the electrocatalytic
advantages of the AFC, but introduces a CO2‐tolerant electrolyte.
In general, an AEM is composed of a polymer backbone on which cationic sites are
tethered. These cationic moieties are not carbonate ions with free mobility as in a liquid
electrolyte. Thus carbonate precipitates cannot form in the AMFC. Transport of OH−
ions within the AEM is between the cationic sites in an analogous manner to the H+
ions that are transported between sulfonic acid sites in the PEMFC membrane. The
AMFC shares the advantage of the PEMFC of being a solid‐state device (there is no
liquid electrolyte to leak) and is built up using catalyst and gas diffusion layers (GDLs)
in much the same way as a PEMFC. Furthermore, corrosion of the bipolar plate is less
of an issue and therefore permits the use of thin and easily manufactured hardware.
5 Several terms and acronyms have been proposed for this type of fuel cell. Anion‐exchange membrane
(AEM) is consistent with the use of PEM in the case of proton‐exchange membrane fuel cells and will be
used throughout this book. The reader will also find references in the literature to alkaline‐electrolyte
membrane fuel cell (AEMFC), hydroxide‐exchange polymer membrane fuel cell (HEMFC) and alkaline
proton‐exchange membrane fuel cell (APEMFC).
Alkaline Fuel Cells
For many years, polymeric anion‐exchange membranes have been employed in seawater desalination plants, the recovery of metal ions from waste waters, electrodialysis
and bio‐separation processes. Unfortunately, however, most of these membranes
possess ionic conductivities that are too low to be considered for AMFC application.
Also, most AEM polymers have poor solubility in the solvents employed in the
production of NafionTM, the membrane that is used in most PEMFCs (see Section 4.2.1,
Chapter 4). The low solubility complicates the fabrication of an AMFC in that, unlike
Nafion, it is more difficult to incorporate an anion‐exchange polymer as a binder in the
electrode layers.
An example of an AEM is that formed by functionalization of a polysulfone via
chloromethylation, followed by reaction with an amine (quaternization) or phosphine
to yield a quaternary ammonium (QA) or phosphonium salt. The salt form of the
membrane can then be treated with KOH to yield a hydroxide‐ion‐conducting AEM in
much the same way that the sodium form of a PEM membrane (e.g., Nafion) can be
treated with sulfuric acid to yield a proton‐conducting membrane. The synthesis
reactions involved in the production of a commercial polysulfone (Udel® from Solvay
Advanced Polymers LLC) are summarized in Figure 5.7.
Membranes prepared by QA chemistry have been the most studied for fuel‐cell
applications, and they have reasonable stability in alkaline environments (especially
CH3
O
O
S
CH3
Chloromethylation
O
CH3
O
S
CH3
O
n
O
CH2Cl
(CH3)3N
Quaternization
O
CH3
O
S
CH3
–Cl+(H C) NH C
3 3
2
O
n
O
CH2N(CH3)3+Cl–
1M KOH
Ion exchange
O
CH3
O
–OH+(H C) NH C
3 3
2
n
O
(CH3)3SiCl + (CH2O)n + SnCl4
CIH2C
O
CH3
CH2N(CH3)3+OH–
S
O
O
n
Figure 5.7 Chemical reaction steps to convert a polysulfone into an anion‐exchange membrane
polymer.
145
146
Fuel Cell Systems Explained
membranes that contain benzyltrimethyl ammonium exchange sites). The general
issues with such AEMs are as follows6:
●
●
●
The diffusion coefficient and mobilities of OH− anions are typically one‐third to one‐
half less than those of H+ in most media, and QA ionic groups are less dissociated
than the typical sulfonic acid groups. Thus, there were concerns that AMEs would
not possess intrinsic ionic conductivities high enough for application in fuel cells.
The OH− ions are effective nucleophiles that potentially cause degradation of the
polymer via (i) a direct nucleophilic displacement and/or (ii) a Hofman elimination
reaction when a β‐hydrogen is present and possibly (iii) a mechanism that involves an
ylide intermediate.7
The AEM must have the chemical stability to withstand the final step in the preparation, i.e., typically, the exchange of chloride (Cl−) ions with OH− ions in a strongly
alkaline solution of NaOH or KOH.
All of the polymer degradation mechanisms are enhanced at high temperatures, and
therefore most AMFC developers are targeting operation at room temperature. Various
starting materials are under investigation for synthesizing anion‐conducting polymers.
Examples are polybenzimidazole (PBI), poly‐ether ketones, polyphenylene oxides and
polyvinyl alcohol grafted with 2,3‐epoxypropyltrimethylammonium chloride. A range
of quaternizing agents are also being evaluated.
Synthetic routes to AEMs other than by quaternization are under investigation, and,
as with PEMFC membranes, there are several well‐practiced methodologies for the
preparation of membranes.8 The polymer can be synthesized directly from a functionalized monomer, polymerized from a monomer with subsequent functionalization or
prepared by functionalizing a commercial polymer. A body of literature is emerging
from which some general remarks concerning AEMs can be made as follows.
Fluorine‐containing polymers generally show higher thermal stabilities than
hydrocarbon polymers. Irradiation of polymer films using X‐rays, γ‐rays or electron
beams is a flexible way to introduce functional groups, but the easiest synthesis route is
to dope inert polymers directly with concentrated KOH solution. For instance, polar
polymers (e.g., polyethylene oxide) can be doped with alkali hydroxides (e.g., KOH), or
ammonium hydroxides such as tetrabutyl ammonium hydroxide. Polybenzimidazole
doped with KOH shows a very high ionic conductivity compared with proton
conductivity in Nafion but could suffer from carbonate precipitation as experienced in
conventional AFC electrolyte solutions.
The development of AMFCs is in its infancy. Single‐cell AMFCs have been built
and tested in the laboratory, but to date no stack demonstrations at the kW scale have
been built.
6 Slade, RCT, Kizewski JP, Poynton, SD and Varcoe JR, 2013, Alkaline membrane fuel cells, in Meyers, RA
(ed.), Encyclopedia of Sustainability Science and Technology, Springer Science + Business Media, New York.
7 Chempath, S, Einsla, BR, Pratt, LR, Macomber, CS, Boncella, JM, Rau, JA and Pivovar, BS, 2008,
Mechanism of tetra‐alkyl ammonium head group degradation in alkaline fuel cell membranes, Journal of
Physical Chemistry C vol. 1123, pp. 3179–3182.
8 Couture, G, Alaaddine, B, Boscheti, F and Amedur, B, 2011, Polymeric materials as anion‐exchange
membranes for alkaline fuel cells, Progress in Polymer Science, vol. 36, pp. 1521–1557.
Alkaline Fuel Cells
5.3
Electrodes
As discussed earlier, although AFCs can be operated over a wide range of temperatures
and pressures, the extent of their applications is quite restricted. Accordingly, there
is no standard type of electrode for the AFC, and different approaches are taken as
determined by performance requirements, operating temperature and pressure and
cost limits. Different catalysts can be used, but this does not necessarily affect the
electrode structure. For example, a platinum catalyst is effective with any of the main
electrode structures that are described here.
5.3.1
Sintered Nickel Powder
When F. T. Bacon designed his pioneering fuel cells in the 1940s and 1950s, he
opted for nickel‐based electrodes in the belief that the expensive platinum‐group
electrocatalysts would never become commercially viable. His electrodes were made
porous through fabrication from powdered nickel, which was then sintered to make
a rigid structure. To enable a good three‐phase contact between the reactant gas, the
electrolyte solution and the solid electrode, the nickel electrode was made in two
layers from two sizes of nickel powder. The procedure gave a wetted, fine‐pore
structure for the liquid side and more open pores for the gas side. Very good results
were achieved, though careful control of the differential pressure between the gas and
the electrolyte solution was necessary to ensure that the liquid gas boundary was
anchored to the electrode (note: wet‐proofing materials, such as polytetrafluoroethylene (PTFE), were not available at that time). This electrode structure was also
selected for the fuel cells employed in Apollo missions. In both the Bacon and Apollo
cells, the anodes employed plain nickel powders, whereas the nickel oxide cathodes
were treated with lithium salts to generate LiNO2 on the surface to provide chemical
stability.
5.3.2
Raney Metals
An alternative method for obtaining a very active and porous form of a metal is the
use of Raney metals; it has been a common practice for AFCs from the 1960s to the
present. The metals are prepared by mixing the required active metal (e.g., nickel)
with an inactive metal, usually aluminium. The mixing is performed in such a way
that distinct regions of aluminium and the host metal are maintained, i.e., the
material is not a true alloy. The mixture is then treated with a strong alkali that
dissolves out the aluminium to leave a porous product with a very high surface area.
The process gives scope for changing the pore size by altering the proportions of the
two metals and by adding small amounts of other metals such as chromium, molybdenum or zinc.
Raney nickel electrodes were employed in many of the demonstrations of fuel cells
that were reviewed in the opening (Section 5.1). Often Raney nickel was chosen for the
anode and silver for the cathode. In the early 1990s, this combination of electrodes was
a feature of the AFC built by Siemens for service in submarines. Raney metals have also
been used as catalyst in a ground‐up form, for the rolled electrodes that are described
in the following section.
147
148
Fuel Cell Systems Explained
5.3.3
Rolled Carbon
Most modern AFCs employ carbon electrodes that are similar to those used in the
PEMFC. In the late 1950s, Karl Kordesch carried out the initial development of carbon
electrodes while he was employed by the Union Carbide Corporation (UCC). The first
UCC electrodes were built up of several layers of carbon black, PTFE and pore‐forming
additives. Catalyst metals included not only nickel but also silver and cobalt. Cell
voltages of around 0.6 V were achieved in air at current densities up to 200 mA cm2. The
work of Kordesch culminated in the demonstration of a hydrazine‐fuelled AFC that
powered a converted Austin A40 van. The vehicle is now in the London Science
Museum.
The latest carbon electrodes now invariably employ carbon‐supported catalyst
metals mixed with PTFE, which are then rolled onto a material such as nickel mesh.
The PTFE acts as a binder, and its hydrophobic properties prevent flooding of the
electrode and provide for controlled permeation of the electrode by the electrolyte
solution. A thin layer of PTFE will often be placed over the surface of the electrode for
two reasons: (i) to control further the porosity and (ii) to impede the electrolyte
solution from passing through the electrode, without the need to pressurize the
reactant gases, a requirement that is necessary with porous metal electrodes. Carbon
fibre is sometimes added to the mix to increase the strength, conductivity and porosity
of the resulting electrode.
Modified papermaking machines can be used to manufacture rolled electrodes at
quite low cost. Such electrodes find application not only in fuel cells but also in metal–
air batteries, for which the cathode reaction is much the same as for an alkali fuel cell.
For example, the same electrode can act as the cathode in a zinc–air battery (e.g., for
hearing aids) and an aluminium–air battery (e.g., to provide reserve power for telecommunications). Such an electrode is shown in Figure 5.8. The carbon‐supported catalyst
is of the same structure as that presented in idealized form in Figure 4.11, Chapter 4.
The catalyst may not always be platinum. Manganese, for example, is an effective
Figure 5.8 The structure of a rolled AFC electrode. The catalyst is mixed with a PTFE binder and rolled
onto nickel mesh. The thin layer of PTFE on the gas side is shown partially rolled back.
Alkaline Fuel Cells
cathode catalyst in both metal–air batteries and AFCs. Commercial rolled electrodes
with a non‐platinum catalyst are readily available at about US$0.01 per cm2 or around
US$10 per ft2, i.e., at a cost that is very low compared with other fuel‐cell materials.
Adding a platinum catalyst increases the cost in line with the loading, but it might only
be by a factor of about three, which, with respect to fuel cells, still gives a very inexpensive
electrode. There are, however, problems elsewhere.
One issue is that because the electrode is covered with a layer of PTFE, the surface is
non‐conductive, and thus a bipolar plate cannot be employed for cell interconnection.
Instead, the cells are normally edge connected. Fortunately, this is not too much of a
constraint given that the nickel mesh running right through the electrode results in a
higher than normal conductivity across the plane of the electrode and thereby renders
edge connection to be a practical option. Edge connection gives a certain flexibility to
stack design in that it is not necessary to connect the positive of one cell to the negative
of the adjacent cell, as must occur with bipolar plates. Instead, series–parallel electrical
connections can be made and inherently improve the performance of the cells by reducing
internal current losses.
The problem of internal shunt currents within an AFC stack is a unique feature of
using a liquid electrolyte that is circulated throughout the stack. The ion‐conducting
electrolyte is in contact with all cells within the stack and can therefore provide an ionic
current pathway between adjacent cells. The path is short if the cells are configured
electrically in series, using conventional bipolar plates between each cell. Note that this
problem does not exist in an MCFC because the current collectors and flow‐field plates
separate the electrolyte of each cell. In the AFC, however, there is no such separation
between cells. By electrically connecting AFCs in parallel, the current path between
cells is elongated so that any loss of voltage caused by movement of electrolyte between
cells is minimized. In practice AFCs may be joined together with a mixture of series and
parallel connections to minimize the losses due to electrolyte circulating between cells.
Apart from the serious problem that crystals of carbonate can form in the pores of the
electrodes from CO2 in the fuel or oxidant gases, there has also been a suggestion that
some carbon dissolution can occur in the AFC catalysts, as happens with PEMFC
carbon cathodes. Extensive studies,9 indicated that the operational life of air electrodes
(PTFE‐bonded carbon electrodes on porous nickel substrates) with CO2‐containing air
at 65°C ranged from 1600 to 3400 h at a current density of 65 mA cm−2, compared with
4000–5500 h when using CO2‐free air under similar conditions. The current density
was not particularly high in these tests, and lifetime was less at higher currents. It was
also found that lower temperatures shorten life, presumably due to a decrease in the
solubility of the carbonate. Note that a lifetime of 3400 h is only 142 days and implies
that such electrodes are only suitable for a limited number of applications.
Gulzow10 describes an anode based on granules of Raney nickel mixed with PTFE
that is rolled onto a metal net in much the same way as the PTFE/carbon‐supported
catalyst. A cathode was prepared likewise, only using silver instead of nickel. It was
claimed that such electrodes are not degraded by CO2.
9 Kordesch, K, Gsellmann, J and Kraetschmer, B, 1983, Studies of the performance and life‐limiting
processes in alkaline fuel cell electrodes, Power Sources, vol. 9, p. 379, ed. By Thompson, J, Academic Press,
New York.
10 Gulzow, E, 1996 Alkaline fuel cells: a critical view, Journal of Power Sources, vol. 61, pp. 99–104.
149
150
Fuel Cell Systems Explained
5.3.4
Catalysts
The AFC stacks in the Orbiter had very high loadings of noble metal catalyst in the
electrodes: a 80 wt.% Pt + 20 wt.% Pd anode catalyst was loaded at 10 mg cm−2 and a
90 wt.% Au + 10 wt.% Pt cathode catalyst was loaded at 20 mg cm−2, each on a silver‐
plated nickel screen. Both catalysts were bonded with PTFE to achieve high performance
at 85–95°C.
The aggressive nature of the alkaline electrolyte is much less than that of acid in
PAFCs or PEMFCs and thereby enables the selection of a broader range of catalysts.
Very high surface area (Raney) nickel can be used at the cathode instead of platinum.
The nickel can, in turn, be enhanced by a catalyst that consists of high surface area
active carbon doped with silver, and iron (or cobalt) macrocyclics such as heat‐treated
cobalt tetra‐phenoxymethyl porphyrins on carbon. As with PEMFC cathode catalysts,
the selection of porphyrins for the oxygen reduction reaction (ORR) has been stimulated
by knowledge of compounds that are involved in the reduction of oxygen in biological
systems.
By raising the temperature, most developers of stationary AFC systems since the
1960s have opted for ‘classic’ non‐noble metals for the catalysts (Raney nickel for the
anode, silver and/or manganese dioxide for the cathode). With a nickel anode catalyst,
activation overpotential is dominant at low current densities, whereas transport processes significantly increase the overpotential at very high current densities. Therefore,
as with the PAFC, it is essential for the AFC to operate within these limits. Unlike the
platinum anode catalyst in the PEMFC, the nickel catalyst in the AFC or PAFC can
undergo permanent oxidation if the current density is allowed to go too high. To avoid
such issues, researchers have been exploring other ion‐conducting materials as catalysts,
specifically spinels and perovskites that are able to tolerate cycling between oxidizing
and reducing conditions.
The cathode catalyst of the AFC is of particular interest since the overpotential at
this electrode contributes most to the voltage loss in the cell. Silver has the highest
electrical conductivity of any element and is approximately fifty times less expensive
than platinum. Moreover, silver is one of the most active catalysts for the ORR — the
metal is competitive to platinum in highly concentrated alkaline media, as well as on
a cost/performance basis. Incidentally, it is interesting to note that cathodes loaded
with silver have also given a longer lifetime (3 years) in alkaline electrolyzers than
platinum‐based cathodes (1 year) under practical chlor‐alkali electrolysis conditions.
The impregnation of silver into a carbon support via the in situ reduction of silver
nitrate (AgNO3) has been shown to produce very fine particles that constitute a high
surface area catalyst for optimum cathode performance. Research has shown11 that
carbonate deactivation can be avoided by replacing the porous carbon support with
porous silver of the form used in commercial water purification membranes. The
catalytic activity of the silver electrode for the ORR can be improved by including a
platinum or manganese dioxide (MnO2) catalyst. Gas accessibility through the silver
is also enhanced by impregnating the pores near the gas surface with Teflon AF
(a microporous form of PTFE).
11 Bidault, F, Kucernak, A, 2011, Cathode development for alkaline fuel cells based on a porous silver
membrane, Journal of Power Sources, vol. 196(11), pp. 4950–4956.
Alkaline Fuel Cells
5.4
Stack Designs
5.4.1
Monopolar and Bipolar
In a monopolar stack, each cell is connected in series with the next by using an electrically
conducing meal strip or wire — the anode of one cell is linked to the cathode of the next.
The arrangement is illustrated in Figure 5.9a and is necessary in the case of recirculating
electrolyte solution cells where the electrodes are coated with non‐conductive PTFE.
(a)
H2
O2
H2
O2
H2
O2
+
–
(b)
Seal
Membrane–electrode assembly (MEA)
Bipolar plate
+
–
H2 O2
H2 O2
H2 O2
H2 O2
Figure 5.9 AFC stack configurations: (a) monopolar and (b) bipolar.
151
152
Fuel Cell Systems Explained
Bipolar AFC stacks are similar in configuration to most PEMFC stacks, as shown
in Figure 5.9b. The bipolar arrangement is more suitable for fuel cells with static
electrolyte solutions where there is no prospect of short circuits occurring because the
electrolyte is held in a matrix material that separates the two electrodes. The downside
of this design is that the matrix has to be relatively thick, which increases the ohmic
loss compared with the relatively thin liquid electrolyte film that can be employed in a
cell with recirculating electrolyte solution. If successful AEMs can be produced, the
bipolar design will be more attractive again, in that it may be possible to use thinner
electrolyte films.
5.4.2
Other Stack Designs
A variation on the more conventional design of static electrolyte solution cell has been
produced by Hydrocell, a Finnish company. The technology uses a gel electrolyte and
has cylindrical geometry, which requires the cells to be connected externally either
electrically in series and/or in parallel according to the required voltage, rather than in
a bipolar arrangement. Another type of stack is the falling‐film fuel cell developed by
Hoechst AG in Germany. The configuration is the same as that of the cell with the recirculating electrolyte solution shown in Figure 5.2, except that the flow of liquid through
the cell is entirely gravity driven. Therefore, the typical height‐dependent hydrostatic
pressure of a column of liquid does not develop, and consequently, the hydrostatic pressure is the same at the inlet and outlet of the cell. The great advantage of the falling‐film
fuel cell is that the pressure difference between the electrolyte on the front side of the
electrode and the gas on the rear remains constant over the whole area of the vertical
electrode and hence is uniform throughout the cell. The absence of a pressure driving
force leads to a stable three‐phase boundary of electrolyte within the GDLs and thereby
minimizes any potential loss of electrolyte solution, with the result that the gap between
the two electrodes can be made very narrow, typically about 0.5 mm. Cells as large as
0.25 × 1 m have been constructed, which, on account of the thin electrolyte layer, exhibit
very large current densities of up to 2.5 A cm−2.
5.5
Operating Pressure and Temperature
Historically, most AFCs have operated well above ambient pressure and temperature.
These two parameters, together with information about the electrode catalyst, are given
for a selection of important types of AFC in Table 5.1. The choice of operating pressure
is dependent on the system design. In general, cells that employ recirculating electrolyte
solution and the falling‐film cell operate at near‐ambient pressure. For spacecraft fuel
cells that used static electrolyte solutions, higher pressures were more common, namely,
from 300 kPa to more than 1 MPa.
The conductivity of the OH− ions is dependent on the temperature and concentration
of the electrolyte solution. Conductivity increases with temperature. The traditional
AFC can be started below 0°C since the freezing point of the electrolyte solution at
typical concentrations of around 30 wt.% is well below that of water. In fact, the
concentration required to achieve the maximum ionic conductivity only increases
from about 30 wt.% at 0°C to 34 wt.% at 80°C. Similarly, the boiling point of the electrolyte
Alkaline Fuel Cells
Table 5.1 Operating parameters for certain AFCs. The pressure data are approximate as there are
usually small differences between each reactant gas.
Fuel cell
Pressure (kPa)
Temperature (°C)
KOH (wt.%)
Anode catalyst
Cathode catalyst
Bacon
500
200
30
Ni
NiO
Apollo
350
230
75
Ni
NiO
Orbiter
410
93
35
Pt–Pd
Au–Pt
Siemens
220
80
n/a
Ni
Ag
Data from: Warshay, M and Prokopius, PR, 1990, The fuel cell in space: Yesterday today and tomorrow,
Journal of Power Sources, vol. 29, pp. 193–200, and Strasser, K, 1990, The design of alkaline fuel cells,
Journal of Power Sources, vol. 29, pp.149–166.
solution is elevated and thus enables the cell to operate up to 230°C if the electrolyte
concentration is increased to 85 wt.%.
The advantages of higher pressure have been considered in Chapter 2 where it was
shown, in Section 2.5.4, that the open‐circuit voltage, Vr, of a fuel cell is raised when the
pressure increases from P1 to P2 according to the relationship:
V
RT
P
ln 2
4F
P1
(5.9)
The demonstration cell of F. T. Bacon operated around 500 kPa and 200°C. Even these
high pressures, however, would only raise the voltage by about 0.04 V if this ‘Nernstian’
effect was the sole benefit. A rise in pressure (and/or temperature) also increases the
exchange‐current density and thereby reduces the activation overpotential at the
cathode (see Section 3.4, Chapter 3). Consequently, the benefit of increased pressure is
much more than equation (5.9) would predict. For example, the very high pressure gave
the Bacon cell a performance that even today would be considered to be remarkable,
namely, 400 mA cm−2 at 0.85 V or 1 A cm−2 at 0.8 V.
The choices of operating pressure, KOH concentration and catalyst are interrelated.
A good example is the transition from the Bacon cell to the system developed for the
Apollo spacecraft. Although the Bacon cell gave an impressive performance, it was a
heavily engineered design that operated at a very high pressure. To reduce mass for
space applications, the pressure had to be lowered. Consequently, the temperature had
to be raised to maintain the performance at an acceptable level. It was then necessary to
increase the concentration to 75 wt.% KOH; otherwise the electrolyte solution would
have boiled. Unfortunately, increasing the concentration considerably lowers the vapour
pressure, as can be seen from Figure 5.10. At ambient temperature, a 75 wt.% KOH
solution is solid, and therefore it was necessary to provide heaters to start the fuel cell.
In the Orbiter system, the concentration was reduced back to 32 wt.%, and the temperature
was set at 93°C.
In many applications of AFCs, the reactant gases are contained in pressurized or cryogenic storage vessels. In such cases, it is necessary to reduce the supply pressure of
each gas to match the operating conditions of the stack. This requires accurate control
to avoid a large differential pressure between anode and cathode compartments. When
pressurized gases are supplied, there also is a risk that leaks may develop. Apart from
153
154
Fuel Cell Systems Explained
4.0
Water
Pressure/MPa
40 wt.% KOH
solution
3.0
2.0
1.0
80 wt.% KOH
solution
50
200
250
300
Temperature/°C
Figure 5.10 Change in vapour pressure with temperature for different concentrations of KOH
solution.
the waste of gas, leakage could lead to the build‐up of explosive mixtures of hydrogen
and oxygen, especially when the fuel cell is for use in confined spaces such as submarines.
One solution to this problem is to provide an outer envelope for the fuel‐cell stack that
is filled with nitrogen at a higher pressure than that of each of the reactant gases. In a
Siemens submarine system, for example, the hydrogen was supplied at 0.23 MPa and
the oxygen at 0.21 MPa, with the surrounding nitrogen gas at 0.27 MPa. Any leak would
result in a flow of nitrogen into the cells that would reduce the performance but would
prevent an outflow of reactant gas.
In AFCs, there is often a difference in the pressure of the reactant gases and/or in the
vapour pressure of the electrolyte solution. For instance, the hydrogen pressure in the
aforementioned Siemens AFC was slightly higher than that of the oxygen. In the Orbiter
fuel cell, the hydrogen gas was kept at 35 kPa below the oxygen pressure. By contrast, the
gases in the Apollo system were at the same pressure, but both were about 70 kPa above
the vapour pressure of the electrolyte solution. There are no rules governing the setting
of reactant pressure — small differences will be required for a variety of reasons, e.g., to
maintain the boundary of the electrolyte solution and gas in the GDLs.
Raising the temperature actually reduces the open‐circuit voltage of a fuel cell, as
explained in Section 2.3, Chapter 2. In practice, however, the magnitude of this effect is
far exceeded by the reduction in the activation overpotential, especially at the cathode.
As a result, increasing the temperature increases the voltage of an AFC. From a wide
survey of results, it has been concluded12 that below about 60°C there is a very large
benefit to raising the temperature, namely, as much as 4 mV per °C for each cell. At this
rate, increasing the temperature from 30 to 60°C would lift the cell voltage by about
12 Hirschenhofer, JH, Stauffer, DB and Engleman, RR, 1995, Fuel Cells: A Handbook, revision 3, pp. 6‐10 to
6‐15, Business/Technology Books, Orinda, CA.
Alkaline Fuel Cells
0.12 V — a major improvement in the context of fuel cells that operate at about 0.6 V
per cell. There is still a noticeable advantage at higher temperatures, but only in the
region of 0.5 mV per °C. It would appear, therefore, that about 60°C would be a minimum
operating temperature for an AFC. At higher values, the choice would depend strongly
on the power of the cell (and thus any heat losses), the pressure and the effect of the
concentration of the electrolyte solution on the rate of evaporation of water.
5.6
Opportunities and Challenges
The AFC is one of the most efficient energy conversion devices, employing a low‐cost
electrolyte and potentially inexpensive electrodes, and is capable of operating at near‐
ambient temperature and pressure. It could be concluded, therefore, that such attributes
would make the technology attractive for many applications.
Unfortunately, successful exploitation of the AFC for terrestrial applications has been
blocked by the incompatibility between the alkaline electrolyte and CO2. Work carried
out in recent years by AFC Energy and others has shown that modern gas‐diffusion
electrodes have somewhat better tolerance to CO2 than earlier porous metal electrodes.
Nevertheless, challenges in separating product water and in mechanically circulating
the electrolyte solution or constraining it within a matrix have continued to hinder the
development of the AFC compared with the PEMFC, which does not suffer from such
fundamental technical issues.
If robust anionic membranes with high OH− conductivity can be produced easily (i.e.,
at low cost), then perhaps the AMFC can compete with the PEMFC in applications such
as fuel‐cell vehicles. The ability to run at higher temperatures may also see a new type
of AFC competing with PAFCs for stationary power generation.
For an AFC to perform reliably over a long period, it is essential to remove the CO2
from the air. Although this is possible, using processes that are practiced industrially
(e.g., the Benfield process or absorption in aqueous alkanolamine solution) such
procedures would substantially increase the cost, complexity, mass and size of the
system. Ahuja and Green have proposed a novel method13 that would only be feasible
when hydrogen is stored as a liquid. Their method takes advantage of the fact that heat
exchangers are needed to warm the hydrogen and cool the fuel cell. The system is
designed in such a way that the incoming air is cooled in a heat exchanger by the liquid
hydrogen as it vapourizes, thereby freezing out CO2 from the air, which can be separated.
The cold air can then be used to cool the cell, and in doing so its temperature is raised
to that required at the cathode inlet. Alternative methods that have received serious
consideration for CO2 removal have included the utilization of zeolite separation
membranes.
Another possibility for AFC application, and which is actually what Bacon had in
mind when developing his AFC designs in the mid‐20th century, is to incorporate the
cells into a regenerative system. Electricity from renewable sources is used to electrolyze
water, and the fuel cell turns the hydrogen and oxygen so produced back into electricity
13 Ahuja, V & Green, R 1988, Carbon dioxide removal from air for alkaline fuel cells operating with liquid
hydrogen – a synergistic advantage, International Journal of Hydrogen Energy, vol. 23(20), pp. 131–137.
155
156
Fuel Cell Systems Explained
as needed. Of course, other types of fuel cell could be employed in such a system, but
here the disadvantages of the AFC would be largely removed, since both reagents would
be free of CO2.
Further Reading
Arges, CG, Ramani, V and Pintauro, PN, 2010, Anion exchange membrane fuel cells,
The Electrochemical Society Interface, vol. 19, pp. 31–35.
Kordesch, KV, 1971, Hydrogen‐air/lead battery hybrid system for vehicle propulsion,
Journal of the Electrochemical Society, vol.118(5), pp. 812–817.
Kordesch, KV and Cifrain, M, 2004, Advances, aging mechanism and lifetime in AFCs with
circulating electrolytes, Journal of Power Sources, vol. 127, pp. 234–242.
McLean, GF, Niet, T, Prince‐Richard, S and Djilali, N, 2002, An assessment of alkaline fuel
cell technology, International Journal of Hydrogen Energy, vol. 27(5), pp. 507–526.
Mulder, G, 2009, Fuel cells –alkaline fuel cells, in Garche, J, Dyer, CK, Moseley, PT, Ogumi,
Z, Rand, DAJ and Scrosati, B (eds.), Encyclopedia of Electrochemical Power Sources,
pp. 321–328. Elsevier, Amsterdam.
157
6
Direct Liquid Fuel Cells
A direct liquid fuel cell (DLFC) generates electricity via the oxidation of a liquid fuel that
requires no preliminary preparation. Most DLFCs use a proton‐exchange membrane
(PEM) as the electrolyte and are therefore closely related to the proton‐exchange
membrane fuel cell (PEMFC). The direct methanol fuel cell (DMFC) is the most mature
version of this technology and is therefore described in the opening section of this chapter.
The cell is available commercially for some low‐power applications; for example, over
35 000 battery chargers employing DMFC systems have been produced by SFG Energy
AG, under the trademark of Energy for You (EFOY).
The remainder of the chapter is devoted to types of low‐temperature fuel cells that
run on alternative fuels that are liquids under normal conditions. Potential candidates
include many alcohols (e.g., ethanol, propanol, propan‐2‐ol) and other organic liquids
(e.g., ethylene glycol, acetaldehyde, formic acid). Some characteristics of these fuels are
given in Table 6.1.
Sodium borohydride (as a solution), which has already been discussed in Chapter 5,
will be also considered as an inorganic fuel for a DLFC.
6.1
Direct Methanol Fuel Cells
Methanol (CH3OH) is a simple alcohol that is liquid at normal temperatures and
pressures (boiling point 64.7°C) and is miscible with water. It is readily available but
has a specific energy (Wh kg−1) that is only half that of gasoline. Nonetheless, in the
early 1990s, methanol was proposed for fuel‐cell vehicles given that it is relatively easy
to reform directly into hydrogen. For instance, Daimler built a demonstration car — the
Necar 3 — that employed an on‐board methanol reformer to generate hydrogen that
fed a PEMFC. If methanol can serve as a fuel, then all the problems associated with
storing hydrogen in a vehicle are swept aside. As has been mentioned already at
various points in Chapters 2 and 3, methanol can, in principle, also be used directly in
fuel cells. The DMFC, in which the methanol is oxidized directly at the anode, has the
advantage of not requiring a fuel processor to convert the methanol to hydrogen.
Consequently, the DMFC could potentially be attractive for small portable systems
where weight can be an issue.
The DMFC was pioneered by Shell Research Limited in England and Exxon‐Alsthom
in France during the 1960s and 1970s. Shell chose a sulfuric acid electrolyte, while
Fuel Cell Systems Explained, Third Edition. Andrew L. Dicks and David A. J. Rand.
© 2018 John Wiley & Sons Ltd. Published 2018 by John Wiley & Sons Ltd.
158
Fuel Cell Systems Explained
Table 6.1 Thermodynamic characteristics of PEMFC and some DLFCs at 25°C and 101.325 kPa.
Fuel cell
Fuel
Weight
(g mol−1)
Number of
electrons
involved
Standard
Theoretical energy
cell voltage density
(V)
(Wh mL−1)
Maximum
efficiency
(%)
PEMFC
Hydrogen
2.01
2
1.23
1.55 (at 70 MPa)
83
DMFC
Methanol
32.04
6
1.21
4.33
97
DEFC
Ethanol
46.07
12
1.15
5.80
97
DEGFC
Ethylene
glycol
62.07
10
1.15
5.85
99
DFAFC
Formic acid
46.03
2
1.41
1.88
106
DPFC
Propanol
DPFC(2) Propan‐2‐ol
60.1
18
1.13
7.35
97
60.1
18
1.12
7.10
97
DEFC, direct ethanol fuel cell; DEGFC, direct ethylene glycol fuel cell; DFAFC, direct formic acid fuel cell;
DMFC, direct methanol fuel cell; DPFC, direct propanol fuel cell and DPFC(2), direct propan‐2‐old fuel cell.
Exxon‐Alsthom pursued an alkaline approach. Despite some good work with alkaline
and buffer electrolyte technology, Exxon terminated its research programme in the
late 1970s. The Shell research continued until the early 1980s when, as a result of the
curtailed growth in oil consumption that resulted from the conservation measures
taken after the 1973 crisis, it became clear that the fears of an imminent oil shortage
were unfounded. The drop in oil prices had pushed the target cost for the DMFC out of
reach. Nevertheless, substantial progress had been made by teams at the Shell Thornton
Research Centre in Chester, United Kingdom, and at the Koninklijke Shell Laboratorium
in Amsterdam, the Netherlands. During the period 1973−1981, the UK effort led to an
improvement in the performance of the fuel electrode by over two orders of magnitude,
and a more detailed understanding of the mechanism of the methanol oxidation
reaction emerged. This outcome is discussed in Section 6.2.2. At the same time, the
Amsterdam laboratory made considerable progress in the development of stable, active,
non‐noble metal catalysts for the air electrode. Some of this work will be addressed
later. The DMFC attracted little attention during the 1980s until the PEMFC emerged
as viable technology at the end of the decade. At this point, several university research
groups, particularly in the United States, started to conduct investigations of DMFCs
based on PEMs. The work carried out by these groups provided the basis for the evolution
of current DMFC technology.
The net specific energy of methanol is higher than other means of storing hydrogen,
particularly as compressed gas or as metal hydride, as indicated in Table 6.2. In general,
liquid fuels have much higher specific energies than gases, and this is an important
factor for a fuel‐cell system destined for transport applications if a long driving range is
to be achieved on a single tank of fuel. Other advantages of the DMFC are ease of
handling of methanol, rapid refuelling and simplicity of design.
A negative aspect of the DMFC is that the oxidation of methanol at the anode is a
much slower reaction than the oxidation of hydrogen at the anode of a PEMFC, as
explained in Section 6.2. Consequently, a DMFC has reduced power output compared
with a PEMFC of similar size and using the same membrane-electrode assembly (MEA).
Direct Liquid Fuel Cells
Table 6.2 Comparison of specific energy (LHV) for common energy storage materials and the most
important hydrogen storage technologies.
Storage method
H2 at 30 MPa in composite
cylinders
H2 in metal hydride cylinders
Specific energy
of fuel
119.9 MJ kg−1
Storage
efficiencya (%)
0.6
33.3 kWh kg−1
119.9 MJ kg
−1
H2 from methanol — ‘indirect
methanol’b
Methanol in strong plastic tanks
for direct use as fuel
Ethanol
119.9 MJ kg
0.65
19.9 MJ kg
6.9
24 MJ kg
95
Gasoline
46.4 MJ kg
95
Diesel
48 MJ kg
13.33 kWh kg−1
22.8 MJ kg−1
6.34 kWh kg−1
95
12.0 kWh kg−1
−1
18.9 MJ kg−1
5.26 kWh kg−1
6.67 kWh kg−1
−1
8.27 MJ kg−1
2.3 kWh kg−1
c
5.54 kWh kg−1
−1
0.78 MJ kg−1
0.22 kWh kg−1
33.3 kWh kg−1
−1
0.72 MJ kg−1
0.20 kWh kg−1
33.3 kWh kg−1
−1
Net specific
energy
44.27 MJ kg−1
11.4 kWh kg−1
95
45.6 MJ kg−1
12.66 kWh kg−1
a
Storage efficiency is defined here as the weight of hydrogen stored per kg of total system. For example, a
compressed gas cylinder that weighs 500 g will store 0.06 × 500 = 30 g of hydrogen, whereas as vessel
containing hydride and weighing the same will contain 32.5 g of hydrogen.
b
An estimated mass of the reformer is included in the case of ‘indirect methanol’, where it is chemically
reacted to produce hydrogen.
c
The storage of these liquid fuels is assumed to be 95% efficient, i.e., the mass of the liquid is 95% of the
total mass of the liquid and storage vessel.
A state‐of‐the‐art DMFC operating at about 50°C with a cell voltage of 0.4 V will produce
around 5 mW cm−2. Raising the temperature to 70–80°C can lead to power densities of
80–100 mW cm−2. Despite these relatively low figures, DMFCs are attractive for some
stationary applications of the small‐to‐medium scale, i.e., up to about 5 kW.
Fuel crossover is another issue with the DMFC and has been discussed briefly in
Section 3.5, Chapter 3. The phenomenon is particularly acute in the DMFC if the
electrolyte is a perfluorosulfonic acid (PFSA) membrane, as described for PEMFCs in
Section 4.2.1, Chapter 4. The water that provides the proton conductivity pathway in
PFSA membranes can readily absorb methanol, and, as a result, the methanol can
quickly migrate from the anode to the cathode. Such action reduces the open‐circuit
voltage of the cell that, in turn, adversely affects the performance of the fuel cell at all
currents. A comparison of the performance of a state‐of‐the art DMFC with that of a
PEMFC is given in Figure 6.1. The shape of the two graphs is broadly similar, but the
voltages and current densities of the DMFC are considerably lower.
Electrolytes for the DMFC and the problem of fuel crossover are further discussed in
Section 6.3. The implications in terms of potential applications for the DMFC are
covered in Section 6.4.
159
Fuel Cell Systems Explained
1.2
A typical performance curve
of H2-PEMFCs
1.0
Cell voltage/V
160
0.8
0.6
A typical performance curve
of direct methanol fuel cells
0.4
0.2
0
100
200
300
400
Current density/mA cm–2
Figure 6.1 Voltage versus current density performance of a 2010 state‐of‐the‐art DMFC and a typical
PEMFC when operating under ambient conditions.
6.1.1
Principles of Operation
The overall reaction in the DMFC can be expressed as follows:
2CH3 OH 3O2
4 H2 O 2CO2
(6.1)
As noted in Section 2.2, Chapter 2, the change in standard Gibbs free energy, g f , for
this reaction is −698.2 kJ mol−1. Six electrons are transferred for each molecule of
methanol that is consumed, and, from equation (2.11), the reversible cell voltage is
therefore given by:
Vr
gf
zF
698.2 1000
1.21 V
6 96 485
(6.2)
The practical voltages obtained are considerably less than this, and the losses are
greater than those for other types of fuel cell. Indeed, one feature that sets apart the
DMFC is that there is considerable voltage loss at both the anode and the cathode.
The anode reaction of the DMFC is discussed in more detail in the following
section.
6.1.2
Electrode Reactions with a Proton‐Exchange Membrane Electrolyte
Methanol can be used as a fuel for both the PEMFC and the alkaline fuel cell (AFC) as
described in Chapters and 5, respectively. For the DMFC with a PEM electrolyte, the
overall anode reaction is:
CH3 OH H2 O
CO2
6H
6e
(6.3)
Direct Liquid Fuel Cells
CH3OH
HCHOH
CHOH
CHOH
1
1
HCHO
CHO
CO
2
2
HCOOH
3
COOH
3
CO2
Figure 6.2 Stepwise reaction paths for oxidation of methanol at a DMFC anode.
The H+ ions move through the electrolyte and the electrons travel round the external
circuit. Note that water is required at the anode, though it is produced more rapidly at
the cathode via the accompanying reaction:
1½O2 + 6H + + 6e − → 3H2 O
(6.4)
Unlike the direct electrochemical oxidation of hydrogen in a PEMFC, reaction (6.3) takes
place in several steps that can take a variety of routes. The first step is the dissociative
adsorption of methanol on the platinum (Pt) catalyst with the release of six protons and
six electrons that gives rise to a large electric current. The product of the dissociation is
a methanolic residue that remains on the catalyst surface, the precise composition of
which is still debated. This surface residue is slowly oxidized to CO2 via reaction with
water or other adsorbed oxygenated species. Despite a plethora of research carried out
by Shell and others before 19801 and by several university research groups since then,
the true mechanism of methanol electro‐oxidation has still to be resolved. What can
generally be agreed is that following the initial dissociative adsorption step,
dehydrogenation involves reaction of the adsorbed species with adsorbed OH groups.
The chart in Figure 6.2 is an attempt to illustrate the steps and possible reaction routes
that may take place during methanol electro‐oxidation. At the top left of the diagram is
methanol; at the bottom right is the main reaction product — carbon dioxide. The lateral
steps from left to right involve ‘hydrogen stripping’ or dehydrogenation, i.e., the removal
of a hydrogen atom and the generation of a proton (H+) and electron (e−) pair. The downward‐
moving steps not only involve the removal of a hydrogen atom and the generation of a
proton–electron pair but also include the addition or destruction of an OH group.
1 Hampson, NA, Willars, MJ, McNicol, BD, 1979, The methanol‐air fuel cell: A selective review of
methanol oxidation mechanisms at platinum electrodes in acid electrolytes, Journal of Power Sources,
vol. 4(3), pp. 191–201.
161
162
Fuel Cell Systems Explained
Any reaction route through the compounds shown in Figure 6.2 from top left to
bottom right is possible, and all have the same net result, namely, the oxidation of
methanol to carbon dioxide and six proton–electron pairs. The compounds connected
by the red arrows are stable compounds, and moving along this sequence might be
considered a ‘preferred’ route. The route can be divided neatly into three steps. First,
the methanol is converted to methanal (formaldehyde), HCHO, i.e.,
CH3 OH
HCHO 2H
2e
(6.5)
The methanal then reacts to form methanoic (formic) acid, HCOOH:
HCHO H2 O
HCOOH 2H
2e
(6.6)
Finally, the formic acid is oxidized to carbon dioxide:
HCOOH
CO2
2H
2e
(6.7)
The sum of the reactions (6.5)–(6.7) is the same as (6.3). The fact that oxidation
proceeds via a number of steps leads to the relatively low reaction rates for direct oxidation of methanol. It can also be seen that the formation of carbon monoxide is possible
and thereby influences the choice of catalyst — an issue that is discussed in Section 6.1.4.
It should be noted in passing that either of the two stable intermediate compounds —
formaldehyde or formic acid — could be used as fuels instead of methanol. Their
specific energies would be considerably less, however, as only four or two electrons
would be produced, respectively, for each molecule of these two fuels.
6.1.3
Electrode Reactions with an Alkaline Electrolyte
If an alkaline electrolyte is used for the DMFC, the anode reaction is:
CH3 OH 6OH
CO2
5H2 O 6e
(6.8)
The OH− ions are generated at the cathode by the reduction of oxygen:
1½O2 + 3H2 O + 6e − → 6H +
(6.9)
Unfortunately, the CO2 produced at the anode will react with a hydroxide electrolyte
to form carbonates and therefore will rule out any prospect of a DMFC based on the
conventional AFC. The advent of anion‐exchange membranes has produced some
renewed interest, which is founded largely on the expectation that a more direct anodic
oxidation will lead to lower voltage losses and also enable the employment of low‐cost
catalysts. Reported power densities from such alkaline DMFCs are, however, much
lower than those for equivalent PEM DMFCs (<10 mW cm−2). Direct methanol fuel cells
based on PEM technology therefore continue to be the preferred option.
6.1.4
Anode Catalysts
Unlike the direct oxidation of hydrogen in the PEMFC, the stepwise oxidation of methanol
leads to a considerable activation overpotential at the DMFC anode. Investigations by
Shell Research in the 1960s found that platinum by itself was rapidly poisoned by
Direct Liquid Fuel Cells
adsorbed reaction products and therefore was unsuitable as a catalyst for methanol
oxidation. As a consequence, a wide range of platinum alloys was examined for catalytic
activity.2 The stepwise oxidation reactions suggest that a bimetallic catalyst may be
suitable, each metal promoting the different types of reaction. Significant enhancements
in activity were found for platinum modified with rhenium, ruthenium, tin or titanium.3
In particular, Shell researchers found that both platinum–ruthenium (Pt–Ru) and
platinum–rhodium (Pt–Rh) performed well as catalysts, but the former was favoured.
Electrodes with catalyst loadings of 10 mg cm−2 gave results that encouraged the
construction of stacks. A Pt–Ru catalyst has to this day continued to be preferred for
methanol oxidation, although some enhancements by additives such as tungstophosphoric acid have been reported. Methanol is believed to be converted to carbonyl
species on the Pt (the ‘left‐to‐right horizontal sequences’ in Figure 6.2), and these
species are further oxidized by OH− groups that are absorbed on the Ru.
As in the PEMFC, the anode catalyst is in the form of finely divided metal particles
supported on carbon black and is prepared using the same procedure as that outlined for the
PEMFC in Section 4.3.1, Chapter 4. To suppress the activation overpotential, the catalyst
loading in DMFCs (often 2–10 mg cm−2) is usually much higher than the corresponding
loading in PEMFCs (0.05–0.5 mg cm−2). A small amount of electrolyte ionomer may be
added to the DMFC catalyst layer to encourage the product gas (CO2) to be expelled
quickly while allowing the methanol–water mixture to penetrate the porous structure.
There is scope for improving the DMFC anode catalyst. For example, catalysts
supported on carbon nanotubes (CNTS) have recently been found to exhibit higher
electrochemical activity than those based on conventional amorphous carbons.
Membrane–electrode assemblies with Pt–Ru+CNT anode catalysts have been
demonstrated to yield greater power densities than reference samples that employ a
Vulcan XC‐72 substrate.4 Electronically conducting oxides (such as oxides of titanium,
tin or tungsten) have also been found to enhance the activity of Pt–carbon catalysts.
6.1.5
Cathode Catalysts
The oxygen reduction reaction at the cathode of a DMFC, as expressed by equation
(6.9), is essentially the same as that for the hydrogen fuel cell with acid electrolyte (reaction (1.2), Chapter 1) except that six electrons are transferred per molecule of methanol,
compared with four electrons per molecule of hydrogen.
Consequently, the same catalyst can be employed, i.e., platinum supported on carbon.
Unfortunately, however, platinum also catalyzes the oxidation of methanol, albeit
slowly, and if there is significant crossover from the anode, there will be a substantial
deterioration in cell performance. A wide range of alternative oxygen reduction catalysts
has been evaluated for use in the DMFC (cf., Section 4.3.2, Chapter 4), but no material
has so far proved to be significantly more methanol tolerant than platinum.
2 Andrew, MR and Glazebrook, RW, 1966, in Williams, KR (ed.), An Introduction to Fuel Cells, p. 127,
Elsevier, Amsterdam.
3 McNicol, BD, Rand, DAJ and Williams, KR, 1999, Direct methanol–air fuel cells for road transportation,
Journal of Power Sources, vol. 83, pp. 15–31.
4 Gan, L, Lu, R, Du, H, Li, B and Kang, F, 2009, High loading of Pt–Ru nanocatalysts by pentagon defects
introduced in a bamboo‐shaped carbon nanotube support for high performance anode of direct methanol
fuel cells, Electrochemistry Communications, vol. 11(2), pp. 355–358.
163
164
Fuel Cell Systems Explained
Unlike the PEMFC, it is not necessary to humidify the air supply to the cathode.
Indeed, excess liquid water must be removed at a rate that is sufficient to prevent
flooding of the cathode. The mass-transfer resistance of oxygen increases when liquid
water is retained in the pores of the macroporous gas-diffusion layer (GDL) and
thereby lowers the cathode performance. Excess water may be removed effectively
by adjusting the hydrophobic properties and pore distribution of the GDL. The
procedure involves bonding a hydrophobic, microporous (pore size 100–500 nm)
carbon–polytetrafluoroethylene (PTFE) layer of 10–30 µm thickness onto the GDL.
The small hydrophobic pores result in a low permeability of liquid water — a feature
that hinders the transport of liquid water from the catalyst layer and forces more liquid
water to be transported back to the anode. The outcome is a lower saturation level in
the microporous layer and a higher rate of oxygen transport into the catalyst layer.
6.1.6
System Designs
There are essentially two different approaches to the design of a DMFC system:
‘active’ and ‘passive’. In the case of active systems, as illustrated in Figure 6.3, pumps,
fans and heat-exchangers are used to provide the stack with a controlled supply of
Negative
terminal
Positive
terminal
Carbon
dioxide out
Methanol and
water mixture
Air and water
vapour out
Water separator
and store
Methanol
sensor
Methanol
tank
Anode
Pump or control
valve
Control valve
for water
resupply
system
Cathode
Pump
Electrolyte
Air in
Figure 6.3 The main components of an ‘active’ DMFC system. Not all the components will always be
present. Larger systems may have additional components, such as heat-exchangers in the fuel system
for cooling, air pumps and methanol condensers in the CO2 outlet pipe. Note that the electrode
connections are shown at the edge for simplicity; normally, the current will be taken off the whole
face of the electrode, as in Figure 1.8, Chapter 1.
Direct Liquid Fuel Cells
reactants and to remove waste heat and product water. In practice, pure methanol
has to be supplemented with a supply of water. As shown previously in reaction
(6.1), water is produced within the fuel cell. The product water will evaporate when
air passes over the cathode; indeed it is likely that the rate of water evaporation will
exceed the rate of water production. (Water management is discussed at length for
the PEMFC in Section 4.4, Chapter 4.) In an active DMFC system, water can be
recovered from the cathode exhaust, stored and resupplied as required with fuel
to the anode. Although recycling water does add system complexity, the use of a
dilute solution of methanol as fuel gives two advantages. First, it helps to reduce the
propensity for crossover as considered in Section 6.3. The concentration has to be
about 1 M (ca. 3 wt.%) to limit methanol crossover. It is important therefore to control
the methanol feed rate to ensure that the optimum concentration is maintained,
either in response to a methanol flow sensor in the feed system or in response to the
output power of the cell. The second advantage with a dilute methanol solution is
that water is in contact with the MEA and thus ensures that the membrane is
hydrated, which is of course important for PFSA membranes. It should be noted
that water management in an active DMFC system is less complex than for most
hydrogen PEMFCs.
In a passive DMFC system, the cells are usually supplied with reactants by means of
diffusion, convection, evaporation and capillary forces. There is no forced recirculation
of the methanol–water mixture or water recovery from the cathode, and, therefore,
passive designs are much simpler than their active counterparts. Passive systems
normally operate with lower power densities and are more suitable for small portable
devices. Methanol is usually fed to the cell as a vapour rather than a liquid. As with
active systems, water management remains an important issue, and, since water is not
supplied to the anode, the MEA in a passive system has to be designed in such a way
that sufficient water is transmitted from the cathode to the anode through the membrane itself. It is osmotic drag that moves water from the anode to the cathode in both
the PEMFC and DMFC. In the case of the PEMFC, water dragged from the anode to the
cathode leads to a build‐up at the cathode and thus to a driving force for back-diffusion
to the anode. Due to the high level of water at the anode in normal liquid‐fed DMFCs,
there is no driving force for back-diffusion of water, so the cathode tends to become
flooded with water.
In both designs of DMFC system, the MEA is very similar to that in the PEMFC. It is
made by hot‐pressing the membrane between the anode and cathode catalyst layers
that are each supported by a carbon GDL.
6.1.7
Fuel Crossover
Fuel crossover was considered in general in Section 3.5, Chapter 3, as it occurs to some
extent in all types of fuel cell and with all fuels. The problem is particularly severe with
a DMFC employing a PEM electrolyte. Since methanol mixes very readily with water, it
is able to penetrate the hydrated membrane and therefore migrate from the anode to
the cathode. The cathode uses a platinum catalyst that will oxidize the fuel, although
not so readily as the Pt–Ru catalyst on the anode. The reaction of the fuel at the cathode
is not only a waste of fuel — it will also reduce the cell voltage, for the reasons explained
in Section 3.5, Chapter 3.
165
166
Fuel Cell Systems Explained
The loss of methanol is often quantified as a ‘crossover current’, i.e., the current that
would have been produced had the methanol reacted fully on the anode. The crossover
current ic is defined by:
ic
nFADm
Ci
I xi
(6.10)
m
where I is the discharge current, n is the number of electrons involved, F is the Faraday
constant, A is the electrode area, Dm is the diffusion coefficient of methanol in the PEM,
Ci is the methanol concentration at the anode–PEM interface, δm is the thickness of the
PEM, ξ is the electro‐osmotic coefficient and xi is the mole fraction of methanol in the
solution. The crossover current can be compared with the useful output current i to
give a ‘figure of merit’ for a DMFC that is expressed as the fuel utilization coefficient ηf.
The coefficient gives the ratio of the fuel that is actually reacted on the anode to the total
fuel supplied, namely:
i
f
i ic
(6.11)
Using the techniques described in the following text, it is possible to raise the figure of
merit up to 0.85 or even 0.90, though 0.80 (or 80%) would probably be a more realistic
value to expect.
6.1.8
Mitigating Fuel Crossover: Standard Techniques
There are four principal ways to reduce fuel crossover:
1) The anode catalyst is made as active as possible, within the bounds of reasonable
cost, so that less methanol is available for diffusion through the electrolyte to the
cathode.
2) The fuel feed to the anode is controlled, so that methanol is not present in excess
at low currents. Clearly, the lower the methanol concentration at the anode, the
lower will it be in the electrolyte, and hence at the cathode, as shown in Figure 6.4.
A concentration of 1 M is regarded as optimum for most DMFC applications.
3) Thick PEM electrolytes should be used (i.e., thicker than what is normal for PEMFCs).
These will not only reduce crossover but also increase the cell resistance, and thus a
compromise will need to be sought. For a DMFC, the membrane normally has a
thickness of between 0.15 and 0.20 mm5 compared with between 0.05 and 0.10 mm
for a hydrogen PEMFC.6
4) In addition to its thickness, the composition of the PEM also has an effect. Studies
have shown that the diffusion and water uptake for 1100EW Nafion is about half that
for 1200EW Nafion. (EW is the weight in grammes of Nafion, in terms of molecular
mass, per sulfonic acid group.)
5 For example, Du Pont’s Nafion 117 at 0.18 mm.
6 For example, Du Pont’s Nafion 112 at 0.05 mm.
Direct Liquid Fuel Cells
120
Crossover current/mA cm–2
100
80
1.0 M
60
40
0.5 M
20
0
0.25 M
0
100
200
300
400
500
600
700
800
Useful output current density/mA cm–2
Figure 6.4 Graph showing how the crossover of methanol to the cathode changes with fuel
concentration at the anode and with load current. (Source: Ren, X, Zelanay, P, Thomas, S, Davey, J and
Gottesfeld, S, 2000, Recent advances in direct methanol fuel cells at Los Alamos National Laboratory,
Journal of Power Sources, vol. 86, pp. 111–116.)
Techniques 1 and 2 reduce the likelihood of crossover by promoting reaction of methanol
at the anode. The anode reaction can also be boosted by increasing the current drawn
from the cell.
6.1.9
Mitigating Fuel Crossover: Prospective Techniques
In addition to the (almost) universally applied four standard techniques, there are other
methods under investigation that are more experimental or at a very early stage of
development. Among these efforts are the following:
1) Use of selective (non‐platinum) cathode catalysts. These materials will stop the fuel
from reacting on the cathode and so eliminate the voltage drop due to the mixed
potential that results from the simultaneous reduction of oxygen and oxidation of
methanol. Examples of non‐precious metal catalysts for the PEMFC are given in
Section 4.3.2, Chapter 4. Organic transition metal complexes appear to be suitable to
some extent except they tend to break down at high temperatures and high potentials.
Transition metal chalcogenides have also been shown to be more tolerant to methanol
than platinum. A few inorganic materials have been proposed as suitable substitutes,
including transition metal sulfides (MoxRuySz, MoxRhySz) or chalcogenides
(Ru1−xMoxSeOz). The methanol tolerance mechanism of these materials is due to the
absence of adsorption sites for methanol dehydrogenation. On the other hand, the
catalytic activity of these materials towards the oxygen reduction reaction is still
much lower than that of platinum. In addition, although the methanol that has crossed
over to the cathode may not react on such materials, it will probably just evaporate
and therefore be wasted. This approach therefore does not offer a complete solution.
167
168
Fuel Cell Systems Explained
2) Insertion of a layer in the electrolyte that is porous to protons but less so to methanol.
If such a material could be found, then this would obviously be a solution to the
problem. Some of the ideas being tried in this area include treating the surface of the
Nafion membrane and also coating it with a very thin layer of palladium. The idea of
using bilayer membranes is now well established as is the incorporation of additives
in the PEM that discourages methanol crossover. The problem with such approaches
is that proton conductivity is usually compromised.
3) Development of more conductive PEMs. In general, it appears that a reduction in the
size and amount of water clusters inside the ionomer will lead to reduced methanol
crossover while retaining good proton conductivity. In terms of alternative non‐
fluorinated PEM materials, blends of polybenzimidazole (PBI) with sulfonated
poly(ether ether ketone) (SPEEK) or sulfonated poly(arylene ether sulfone)
(BPSH‐30) or sulfonated poly(arylene ether ether nitrile) (m‐SPAEEN‐60) show the
most promising characteristics for DMFCs.
6.1.10
Methanol Production
The potential of the DMFC, as well as the steam reforming of methanol given consideration in Chapter 8, relies on the fact that methanol is produced in bulk at a reasonable cost. Production in 2013 amounted to 70 million tonnes and global demand in
2016 is over 90 million tonnes. According to the US Methanol Institute, about 30% of
the global demand is for producing formaldehyde (employed in the preparation of
urea–formaldehyde and phenol–formaldehyde resins for particle board and other
construction materials). Methanol is also required for the production of other important
industrial chemicals such as acetic acid, various cleaners and windshield washer fluid
for cars. Only about 2% is currently used as a fuel, typically blended with other
hydrocarbons.
Methanol can be formed by the reaction of hydrogen and carbon monoxide over a
suitable catalyst at high pressures and, therefore, can be obtained by the steam reforming
of natural gas or another hydrocarbon fuel or biofuel (see Chapter 10). As methanol is
required in large quantities for industrial purposes, considerable effort has been
expended on making the production process as efficient as possible. The overall
efficiency is ~70% and ~60% for conversion from natural gas and renewable fuels,
respectively. Such processing efficiencies mean that the cost of methanol is principally
governed by the cost of the raw fuel. Posted prices by Methanex in July 2016 range from
US$240 to 275 per tonne or between US$0.72 and 0.82 per gallon, which puts the fuel
at a price level comparable with that of gasoline.
6.1.11
Methanol Safety and Storage
The use of methanol in consumer products, such as power supplies for portable
electronics equipment, raises potential safety concerns. The first issue is flammability
and the fact that methanol burns with an invisible flame. The same situation, of course,
applies to hydrogen, and studies have shown that hydrogen and methanol are both
equal from a safety point of view and both are better than gasoline.
The second issue relates to the toxicity of methanol. The chemical is a poison, made
worse by the fact that it can mix easily with water. Furthermore, it does not have a taste
Direct Liquid Fuel Cells
that would make it immediately repellent. This ‘drinkability’ problem causes methanol
to be considerably more dangerous in everyday use than other fuels — such as
gasoline — which are also poisonous and in wide circulation. The safety arguments are
fairly complex; however, because methanol is naturally present in the human body it is
perfectly safe in small quantities. Indeed, methanol is produced in the body by the
digestion of a wide range of natural products (e.g., fruit) as well as man‐made additives
such as the aspartame sweetener used in ‘diet’ drinks. By contrast, when liquid methanol
enters the body by direct contact with the skin, by drinking or by inhaling the vapour,
it transfers to the liver where it is converted to formaldehyde. In turn, the formaldehyde
oxidizes to formic acid that interferes with the function of mitochondria and thereby
causes toxic injury to the retina and optic nerve that frequently results in blindness
and, in extreme cases, death. Data collected by the American Association of Pollution
Control Centers suggest that most deaths from methanol poisoning are suicides or the
result of self‐harm actions. Methanol systems must therefore be designed in such a
way that deliberate and premeditated drinking of methanol is very difficult — stopping
accidental consumption is not enough. The usual practice of including an additive with
the liquid that makes it undrinkable could be problematic if the additive interferes with
the operation of the fuel cell. Inhaling methanol vapour is particularly dangerous and
needs to be considered in any system in which methanol is vapourized.
The storage of methanol poses few problems. The liquid should be kept in airtight
stainless steel or plastic containers because it can absorb moisture from the atmosphere, and even in dilute form a methanol–water mixture is quite corrosive. Methanol
can also dissolve some polymers, so care must be taken over the choice of materials
for the storage vessel, gaskets and tubing. For small‐scale or portable applications,
storage concerns have largely been addressed and both the International Civil Aviation
Organization’s Dangerous Goods Panel and the US Department of Transportation
now permit approved fuel cells with installed methanol cartridges to be carried by
passengers and crew on aircraft. Large‐scale methanol storage requires special
measures to be taken in terms of venting and fire safety, but these are beyond the
scope of this book.
6.2
Direct Ethanol Fuel Cells
A direct ethanol fuel cell (DEFC) is more acceptable to consumers and for commercial
markets compared with cells using toxic methanol. Ethanol has a higher specific energy
(6.67 kWh kg−1, LHV) than methanol (5.54 kWh kg−1, LHV) and a higher boiling point
(78°C). It is potentially attractive for both stationary and mobile fuel‐cell systems.
Ethanol can be obtained from renewable plant material such as food crops (e.g., cereal
grains, sugarcane, sugar beets) or other forms of biomass such as switch grass and forest
residue. It can therefore be deemed to be a ‘zero‐carbon fuel’, unlike methanol made
from natural gas, since the CO2 produced when bioethanol is oxidized is returned to the
atmosphere to be consumed by renewable plants. Commercially, ethanol is produced
by three main routes: fermentation of starch and sugar using yeast, processing of
biomass with bacteria and reaction of ethene (which can be produced from oil) with
steam over a catalyst.
169
170
Fuel Cell Systems Explained
6.2.1
Principles of Operation
In the DEFC, ethanol in an anhydrous form or diluted with water, either as liquid or
vapour, is fed directly into the anode, with air supplied to the cathode. Conceptually, as
with the DMFC, the DEFC may employ either acid or alkaline electrolytes. The direct
overall oxidation of ethanol can be represented as:
C 2 H5 OH 3O2
G
2CO2
1325 kJ mol
1
3H2 O
Vr
(6.12)
1.145 V
The standard reversible voltage at 1.145 V is less than that of the DMFC (1.21 V), but
unlike the DMFC in which there are 6 electrons involved in the electrochemical
reactions with ethanol oxidation, there are 12 electrons. In an acid DEFC, the anode
reaction is:
C 2 H5 OH 3O2
2CO2 12H
12e
(6.13)
The corresponding oxygen reduction at the cathode is:
3O2 12e
12H
6CH2 O
(6.14)
As was the case with the DMFC, the earliest work on DEFCs was with an alkaline
electrolyte (aqueous KOH). In this version, hydroxide ions are the electroactive species,
and the electrode reactions are as follows:
Anode : C 2 H5 OH 12OH
Cathode : 3O2 6CH2 O 12e
2CO2
9H2 O 2e
12OH
(6.15)
(6.16)
Unfortunately, as with the DMFC and AFCs in general, there is the problem of contamination of the electrolyte by CO2 that is generated at the anode of the DEFC in addition
to being present in the air supplied to the cathode.
Some work has been conducted in which the aqueous KOH electrolyte is replaced by
an OH− ion‐conducting membrane. To date, published research suggests that with
membranes such as alkali‐doped PBI, the benefits may not be substantially better than
the performance obtained from a DEFC with a proton‐conducting membrane.
6.2.2
Ethanol Oxidation, Catalyst and Reaction Mechanism
The main challenge in the electrochemical oxidation of ethanol is the breaking of the
C─C bond in the molecule. Given that PEM electrolytes function at low temperatures
(below 100°C), the requirements placed on the catalyst are even more demanding than
those for methanol oxidation. Nevertheless, both carbon‐supported Pt–Ru and Pt–Sn
catalysts have been found to be suitable. As with methanol oxidation, bifunctional
catalysts are required because Pt can easily be poisoned by the carbonyl reaction
intermediates. In this respect, the reaction mechanism for ethanol oxidation is more
complex and even more uncertain than that for methanol.
The present understanding of the reaction in an acidic DEFC can be summarized as
follows: The reaction proceeds via a multistep mechanism that involves a number of
Direct Liquid Fuel Cells
CH3
Ethanol
HO
H
H
H2O+
OH2
H
H3C
O
H
OH2
CH3
H3O+
O
1e–
0
Acetaldehyde
O
H
0.8
1e–
Adsorbed
acetyl
CH3
OH2
H3
O+
O
C
Acetaldehyde
EV (RHE)
CH3
O
CH3
H
Acetic acid
OH
OH
1e–
H3O+ + 1e– 2H2O
Adsorbed
acetyl
CH3 H
CH4
E < 0.2 V (RHE)
H 3C
C
–
O
–
–
O
C
H
O
OH2
H2O+
O
C
O
E > 0.5 V (RHE)
1e–
Figure 6.5 Mechanism of ethanol electro‐oxidation at platinum surface in acid medium. (RHE =
reversible hydrogen electrode.) (Source: Reproduced with permission from Vigier, F, Rousseau, S,
Coutancean, C, Leger, J‐M and Lamy, C, 2006, Electrocatalysis for the direct alcohol fuel cell, Topics in
Catalysis, vol. 40(1), pp. 111–121. Reproduced with permission of Springer.)
adsorbed reaction intermediates and by‐products, which emanate from the incomplete
oxidation of ethanol. The major intermediates have been identified as adsorbed carbon
monoxide (CO) and C1 and C2 hydrocarbon residues, whereas acetaldehyde and acetic
acid have been detected as the main by‐products. Information obtained from electrochemical and spectro‐electrochemical studies has led to the proposal and general
acceptance of the reaction mechanism outlined in Figure 6.5.7 The first reaction product from the dissociative adsorption of ethanol on platinum is acetaldehyde, which
requires the transfer of only two electrons per ethanol molecule. Acetaldehyde has to
re‐adsorb on the catalyst to complete its oxidation into either acetic acid (CH3COOH)
or CO2 with methane produced at low potentials. Unlike acetaldehyde, it is difficult to
oxidize acetic acid further at low temperatures, so it becomes a ‘dead-end’ in the reaction.
The possible formation of many intermediate products rather than the complete
conversion of ethanol to CO2 leads to a significantly higher overpotential at the DEFC
7 Vigier, F, Rousseau, S, Coutancean, C, Leger J‐M and Lamy, C, 2006, Electrocatalysis for the direct alcohol
fuel cell, Topics in Catalysis, vol. 40(1), pp. 111–121.
171
172
Fuel Cell Systems Explained
anode in addition to the drying of Nafion® and other perfluorinated sulfonic acid (PSFA)
membranes with concomitant loss in proton conductivity.
A slight improvement in the rate of ethanol oxidation can be achieved by increasing
the operating temperature, and, in 1998, the first reasonable performance given by a
DEFC was reported.8 The cell employed a Pt–Ru anode catalyst and Pt cathode catalyst,
both supported on carbon, together with a composite membrane synthesized from
Nafion and silica (SiO2). Although the power density of 110 mW cm−2 was about half of
that obtainable with a DMFC under the same conditions (0.6 A cm−2 and 0.4 V at 550 kPa
and 145°C), the high selectivity towards the formation of CO2 (95%) enhanced the
prospect of a viable DEFC. Research on different catalyst materials suggests that the
first step of the oxidation reaction occurs on the Pt surface and is not enhanced by
alloying the Pt with other metals. Indeed, the combination of Ru with Pt appears to
inhibit cleavage of the C─C bond. On the other hand, the addition of Ru restricts the
formation of unwanted intermediates and thereby improves the selectivity towards
CO2. A substantial body of literature on ethanol oxidation catalysts has emerged over
the past 20 years and has shown that Pt modified with Sn and/or Ru are effective
combinations. In general, and in contrast to the DMFC, Pt in conjunction with Sn is
presently the more active binary catalyst for ethanol oxidation. Catalysts comprising
Pt–Sn–Ru with a nominal Ru:Sn atomic ratio of less than 1 appear overall to be the most
promising ternary anode catalysts, despite the fact that the Sn and Ru both inhibit
cleavage of the C─C bond.
For a DEFC with an alkaline electrolyte, the anode reaction is subtly different from
that with an acid electrolyte, cf., reactions (6.13) and (6.15), although obviously the
cleavage of the C─C bond is a significant issue in both cells. Interestingly, platinum on
carbon has twice the activity for ethanol oxidation in an AFC than it does in an acid cell.
In early work, non‐precious metal catalysts (iron, nickel and/or cobalt supported on
carbon) showed promise as anode catalysts in alkaline systems, but they were less active
than Pt–C. Palladium, which gives no activity at all in acid fuel cells, has recently been
found to be effective in alkaline DEFCs, especially if combined with certain oxides such
as ceria (CeO2) or titania (TiO2). Unfortunately, most of the catalysts in alkaline systems
do not oxidize the ethanol completely to carbon dioxide but typically stop at the
oxidation level of acetic acid, and this may be acceptable for some applications. Despite
this limitation, NDC Power based in Cheyenne, USA, has developed a platinum‐free
DEFC that has been scaled up to 10‐kW prototype stacks for military applications.
6.2.3
Low‐Temperature Operation: Performance and Challenges
The performance of state‐of‐the‐art DEFCs is significantly inferior to that of DMFCs.
The main challenge is the slow kinetics of electrochemical reactions on both electrodes,
especially at the anode. The high overpotential at this electrode is due to both the
difficulty in breaking the strong C─C bond in ethanol and the slow reaction rate given
the number of reaction steps, coupled with the low selectivity to complete oxidation
and CO2 production. Although bi‐/tri‐metallic Pt‐based catalysts have been extensively
8 Arico, AS, Creti, P, Antonucci, PL, and Antonucci, V, 1998, Comparison of ethanol and methanol
oxidation in a liquid‐feed solid polymer electrolyte fuel cell at high temperature, Electrochemical and
Solid‐State Letters, vol. 1, pp. 66–68.
Direct Liquid Fuel Cells
studied for the anode, the activity of the best catalysts is still too low for practical
application. Compared with the anode, oxygen reduction at the cathode is relatively
fast; nevertheless, there is still room for improvement of the Pt–C catalyst that is the
most often employed. As with the DMFC, another challenge is the crossover of ethanol
from anode to cathode. Not surprisingly, raising the temperature of operation improves
the performance of the DEFC, as has been demonstrated9 for a cell with a composite
Nafion–silica membrane — the peak power density increases from about 60 mW cm−2
at 90°C to 90 mW cm−2 at 130°C.
6.2.4
High‐Temperature Direct Ethanol Fuel Cells
Both molten carbonate (MCFC) and solid oxide (SOFC) fuel cells are able to accept
alcohols directly as fuels. When fed to the anode in either of these two technologies,
ethanol is able to react on the nickel‐containing anode or catalyst via a number of
reactions, such as steam reforming, partial oxidation, autothermal reforming and
dry (CO2) reforming. The reactions are discussed in Chapter 10. Unfortunately,
carbon deposition can be an issue, particularly in high‐temperature SOFCs (>800°C).
Ethanol may decompose and deposit carbon either in the inlet channels of the SOFC
or on the nickel anode material. Unless sufficient steam is present to suppress the
reaction, carbon formation can prove to be a significant problem for both MCFCs
and SOFCs. There are many factors that determine how and where carbon may
occur and include, for example, the temperature of operation and the composition of
the anode catalyst. To reduce the operating temperature for ethanol oxidation, some
work has been undertaken with a combination of SOFC and MCFC electrolytes,
namely, incorporating molten carbonate in a solid oxygen ion‐conducting matrix.
With such material, power densities as high as 500 mW cm−2 have been reported for
ethanol at 580°C.10
6.3
Direct Propanol Fuel Cells
In the search for ways to reduce fuel crossover, alcohols other than methanol and
ethanol have been evaluated on Pt or Pt–Ru electrodes. For example, the electro‐oxidation
of propan‐2‐ol on platinum electrodes was studied by several research groups during
the 1990s. The potential advantages of this fuel are as follows: (i) it is relatively less toxic
than other alcohols, (ii) at low potentials, it is less prone to anode poisoning and (iii) it
has better resistance to crossover and cathode poisoning. The performance of the fuel
cell for a given catalyst system has been associated with the effect of parameters such as
2‐propanol concentration, anode and cathode fuel flow rates, cell temperature and
oxidant back-pressure. Optimization studies have shown that a power density of
45 mW cm−2 can be achieved with 1.5 M propan‐2‐ol at a cell temperature of 80°C when
9 Di Blasi, A, Baglio, V, Stassi, A, D’Urso, C, Antonucci, V and Aricò, AS, 2006, Composite polymer
electrolyte for direct ethanol fuel cell application, ECS Transactions, vol. 3(1), pp. 1317–1323.
10 Mat, DM, Liu, X, Zhu, Z and Zhu, B, 2007, Development of cathodes for methanol and ethanol
fuelled low temperature (300–360°C) solid oxide fuel cells, International Journal of Hydrogen Energy, vol. 32,
pp. 796–801.
173
174
Fuel Cell Systems Explained
using a Nafion 117 PEM and anode and cathode loadings of 4 mg Pt–Ru cm−2 and
1 mg Pt cm−2, respectively.
Direct propan‐2‐ol fuel cells have also been tested with both Nafion and H3PO4‐doped
PBI membrane electrolytes bonded with Pt–Ru catalysts. The initial product of
oxidation at low potentials (<0.4 V vs. RHE) is acetone, which undergoes further oxidation
at higher potentials. Obviously, with alcohols of higher molecular weight, there is the
prospect of more intermediates forming at the anode catalyst that have the potential to
poison or degrade the catalyst. Consequently, it remains to be seen whether systems
employing propanol or other alcohols will fare any better than methanol or ethanol.
6.4
Direct Ethylene Glycol Fuel Cells
Ethylene glycol (CH2OH)2 is another alcohol that is manufactured on a large scale;
globally more than seven million tonnes are produced each year. It is widely available as
antifreeze for the radiators of automobiles and as an ingredient of popular plastics, such
as polyethylene terephthalate (PET). It has a number of features that make it superior to
methanol for fuel cells, namely:
●
●
●
●
High boiling point 198°C (cf., methanol 64.7°C) and low vapour pressure.
High energy capacity 4.8 Ah ml−1 (cf., methanol 4.0 Ah ml−1).
An established distribution infrastructure for the automobile industry.
An OH group on each carbon atom, therefore more easily oxidized than ethanol.
6.4.1
Principles of Operation
The operation of direct ethylene glycol fuel cell (DEGFC) is exactly the same as that of
both the previous direct alcohol fuel cells. With an acid electrolyte, the oxidation of fuel
at the anode can be represented by:
CH2 OH
2
2H 2 O
2CO2 10H
10e
(6.17)
Thus, if ethylene glycol could be oxidized to CO2 completely, 10 electrons would be
obtained from one molecule of the fuel. Even with the most active anode catalyst,
however, complete oxidation is rarely achieved, and, as with ethanol, the energy required
to break the C─C bond is the main barrier. Most DEGFCs have been based on PEM
electrolytes in which protons transfer from the anode to the cathode. When such cells
operate at room temperature and potentials below about 0.9 V, ethylene glycol is
degraded to oxalic acid, which cannot be oxidized further. Electro‐osmotic drag also
transfers water from the anode to the cathode; consequently fuel crossover is also an
issue. A mixed potential can be formed between the oxidation of transferred ethylene
glycol and the reduction of oxygen at the cathode.
In the 1970s, Siemens built an alkaline DEGFC system that employed circulating
KOH solution as the electrolyte. The stack consisted of 52 cells that produced 28 V at
4.5 A (normal power output 125 W) and 16 V at 14 A (225 W peak power). The anode
catalyst was a platinum–palladium–bismuth alloy with a Pt + Pd loading of about
5 mg cm−2. In the alkaline version of the cell, the anode reaction can be expressed as:
Direct Liquid Fuel Cells
CH2 OH
2
10OH
2CO2
8H2 O 10e
(6.18)
As with other alkaline cells, CO2 is formed at the anode and thereby has an immediate
impact on the local pH of the electrolyte. If the liquid electrolyte is replaced by an
anion‐exchange membrane, stability under reduced pH has also proved to be an issue,
and further effort must be undertaken to develop an anionic membrane that is robust
enough for this type of fuel cell.
6.4.2
Ethylene Glycol: Anodic Oxidation
The probable pathway for ethylene glycol oxidation is shown in Figure 6.6.11 As with
other alcohols, the oxidation process involves several consecutive and parallel
adsorption–desorption reaction steps that feature intermediates. The reaction can
progress easily and stop with the formation of oxalic acid or oxalates since the C─C
bond cannot be cleaved. Some C2 products have also been detected in the product
stream, notably formic acid and formaldehyde.
The partial oxidation of ethylene glycol up to oxalic acid (involving the transfer of
eight electrons) yields 3840 Ah L−3 (3450 Ah kg−1); this value is comparable with that of
the complete oxidation of methanol to CO2, namely, 3970 Ah L−3 (5019 Ah kg−1). This is
also the case with ethanol. Partial oxidation to intermediate products is certainly not
desirable — not only because of the loss of energy but also on account of the toxicity of
substances such as glyoxal, glocyoxylic acid and oxalic acid. Raising the platinum
loading on the anode catalyst increases the probability of re‐adsorption of such
intermediates and leads to more complete oxidation. The alternative approach, as
practised with the other direct alcohol fuel cells, is to use platinum alloys such as Pt–Ru
and Pt–Sn as oxidation catalysts. Many bimetallic and tri‐metallic catalysts supported
on traditional amorphous carbons or multi‐walled carbon nanotubes (MWCNTs) have
COOH
2H+
CH2OH
CH2OH
Ethyelene glycol
2H+
2e–
CHO
CH2OH
Glycolic acid
2e–
2e–
2H+
COOH
CHO
Glyoxalic acid
CH2OH
Glycol aldehyde
2e–
2H+
2e–
COOH
COOH
Oxalic acid
2H+
2CO2
2e–
2e–
CHO
2H+
CHO
Glyoxal
2H+
Figure 6.6 Stepwise reaction scheme for the oxidation of ethylene glycol.
11 Ogumi, Z and Miyazaki, K, 2009, Direct ethanol glycol fuel cells, in Garche, J, Dyer, CK, Moseley, PT,
Ogumi, Z, Rand, DAJ and Scrosati, B (eds.), Encyclopedia of Electrochemical Power Sources, pp. 412–419,
Elsevier, Amsterdam.
175
176
Fuel Cell Systems Explained
been tested for ethylene glycol oxidation. The reaction scheme shown in Figure 6.6 is
notably more complex than that for methanol or ethanol, and the ability of the different
catalyst compositions to oxidize the various reaction intermediates is not well established.
Elucidation of both the nature of the reaction pathway and the role of the catalyst
components requires further research.
6.4.3
Cell Performance
Since the boiling point of ethylene glycol is higher than that of either methanol or ethanol
and its vapour pressure is lower, DEGFCs are able to operate at higher temperatures.
As the temperature is raised, so the rates of the electrochemical reactions increase,
together with an enhancement in the proton conductivity of the membrane, but with
the downside of an escalation in crossover. Improvements to the selectivity for oxygen
reduction by the cathode catalyst will therefore be necessary if the advantages of higher
temperatures are to be realized. To date, most of the development of DEGFCs has
focused on the use of PFSA membranes, principally the various forms of Nafion. Few
studies have employed membranes that are capable of operation at temperatures above
about 90°C.
Two other factors have been found to affect the performance of DEGFCs, namely,
pH and ethylene glycol concentration. In an alkaline cell with a KOH electrolyte, the
generation of CO2 at the anode can reduce the pH of the electrode and thereby lower
the cell performance through adsorption of carbonate ions on the platinum catalyst.
This adverse effect can be mitigated by employing bimetallic catalysts (e.g., Pt–Ru). In
alkaline electrolytes, high concentrations of ethylene glycol can lead to increased
crossover, all other factors being equal. At the same time, it has been reported that
increasing the concentration from 1 to 6 M delivers higher current densities.
6.5
Formic Acid Fuel Cells
Formic acid (also known as methanoic acid), HCOOH, is the simplest of the carboxylic
acids with just one carbon atom in the molecule. As noted in the previous sections, it is
formed as an intermediate product in the direct oxidation of other alcohols and can be
used as a fuel in its own right. Formic acid has the lowest energy density among the
liquid alcohols listed earlier in Table 6.1, but this is offset by the higher theoretical cell
voltage (1.41 V) compared with other direct acid fuel cells.
The overall reaction of the direct formic acid fuel cell (DFAFC) can be represented by
HCOOH + ½O2 → CO2 + H2 O
(6.19)
As with the alcohols considered previously, interest in formic acid has been primarily
for the low‐temperature fuel cells, although many other systems could probably operate
with this fuel. Formic acid will react with alkaline electrolytes to produce salts (formates)
that cause cell degradation. This effectively rules out using the fuel with alkaline cells
and leaves PEMFC systems as the preferred option.
The total direct anodic oxidation of formic acid involves the transfer of only two
electrons per molecule, i.e.,
HCOOH
2H
CO2
2e
(6.20)
Direct Liquid Fuel Cells
Unlike the higher alcohols, in which many electrons are transferred in multiple steps,
there are few steps in the electro‐oxidation reaction of formic acid. This feature increases
the likelihood for complete oxidation to CO2 and therefore better fuel utilization.
Note that, unlike methanol (reaction (6.3)) or ethanol (reaction (6.14)), water is not required
for the oxidation of formic acid, and there is no C─C bond to be cleaved. In the case of the
DMFC, for example, methanol concentrations of 0.5 M are required, whereas formic acid
can, in theory, be used in the fuel cell without dilution. Although it is necessary to keep PFSA
membranes hydrated, the complex water management provision that is usually required with
PEMFC‐related fuel cells is not an issue when employing formic acid as the fuel.
6.5.1
Formic Acid: Anodic Oxidation
Formic acid displays higher activity for oxidation than the alcohols discussed earlier,
and therefore both platinum and palladium are effective as anode catalysts in DFACs.
It is now well established that the oxidation of formic acid on a platinum metal surface
takes place through a dual‐pathway mechanism.12 The direct path can be represented
by the fast reaction (6.20), which proceeds via highly reactive intermediates. By contrast,
the oxidation can follow an indirect pathway that involves site blocking or poisoning
intermediates, such as the strongly adsorbed ─COOH intermediate also encountered
during the electro‐oxidation of methanol on platinum, i.e.,
Pt HCOOH
Pt COOH
x
H
(6.21)
e
The adsorbed ─COOH intermediate is then oxidized by adsorbed Pt–OH species,
namely:
Pt COOH
x
Pt OH
2Pt CO2
H2 O
(6.22)
The direct and indirect pathways are manifested in cyclic voltammograms performed
on platinum electrodes, as shown in Figure 6.7. The peak current at 0.5–0.6 V (with
respect to the RHE) corresponds to the direct pathway, whereas the peak approximately
0.9 V corresponds to the indirect pathway. In most practical cases of formic acid
oxidation, the indirect mechanism predominates. As with DMFCs, Pt–Ru is found to
perform well as an anode catalyst, particularly at low temperatures.
With palladium catalysts, the direct pathway is dominant, and no adsorbed reaction
intermediates are detected. The most studied alloy for formic acid is Pd–Pt that gives
a high activity (conversion efficiency of about 50%) which is comparable with that of
hydrogen on a Pt electrode in a PEMFC.
6.5.2
Cell Performance
The solubility of formic acid in water is of a similar order to that of methanol.
Nevertheless, it poses less of a crossover issue with PFSA membranes because it partially
ionizes to create formate ions (HCOO−) that are repelled by the negatively-charged
sulfonate groups on the membrane molecules. Formic acid is much less hygroscopic than
12 Capon, A and Parsons, R, 1973, The oxidation of formic acid on noble metal electrodes III, intermediates
and mechanism on platinum electrodes, Journal of Electroanalytical Chemistry, vol. 45, pp. 205–231.
177
Fuel Cell Systems Explained
150
Current / mA
178
100
50
0.8
0
0.2
1.2
Electrode potential/ V vs. RHE
Figure 6.7 Cyclic voltammogram for a smooth Pt bead electrode (diameter = 0.15 mm) at 25°C in
0.5 M H2SO4 containing 0.1 M HCOOH. Scan rate = 140 mV s−1. (–) continuous reading and (‐‐‐) after
5 min at open circuit. (Source: Reproduced from Capon, A and Parsons, R, 1973, The oxidation of
formic acid on noble metal electrodes III. Intermediates and mechanism on platinum electrodes,
Journal of Electroanalytical Chemistry, vol. 45, pp. 205–231. Reproduced with the permission of
Elsevier.)
methanol, and this also helps to reduce crossover since, even in its most concentrated
form, it will not wet hygroscopic materials such as the membrane and GDLs. The
high‐power density for formic acid in combination with high concentration and low
crossover results in high current density being achieved for DFACs compared with
other liquid‐fuelled cells. For instance, a DFAC with an open‐circuit voltage of 0.72 V
has been shown to deliver a maximum current density of 134 mA cm−2 and power
outputs of up to 49 mW cm−2 with 12 M formic acid.13
6.6
Borohydride Fuel Cells
The possibility of using sodium borohydride as a fuel for AFCs was introduced in
Chapter 5, but there are few published references to borohydride fuel cells before 2000. As
a fuel, sodium borohydride can safely be shipped as a white solid or as a 30 wt.% aqueous
solution. The chemical has wide application in chemistry as a reducing agent for converting
sulfur dioxide to sodium dithionate that serves as a bleaching agent for wood pulp and in
the dye industry. The chemical is also used in the synthesis of vitamin A and for the reduction
of aldehydes to ketones and alcohols in the preparation of various antibiotics.
In acid solution or in the presence of a catalyst, sodium borohydride will react with
water to produce hydrogen:
NaBH 4
2H 2 O
NaBO2
4 H2
(6.23)
13 Rice, C, Ha, S, Masel, RI, Waszczuk, P, Wieckowski, A and Barnard, T, 2002, Direct formic acid fuel
cells, Journal of Power Sources, vol. 111(1), pp. 83–89.
Direct Liquid Fuel Cells
It could therefore act as a source of hydrogen for a conventional PEMFC or AFC. The
term ‘indirect borohydride fuel cell (IBFC)’ is used to define such a fuel cell that is
supplied with hydrogen generated in a separate reactor through the reaction of borohydride and water. By contrast, the ‘direct borohydride fuel cell (DBFC)’ refers to an AFC
in which the borohydride is decomposed and oxidized directly on the anode. In theory,
hydrogen should not be produced in the DBFC by borohydride decomposition and
consequently a higher efficiency may be achieved.
The direct decomposition of sodium borohydride on the anode of a DBFC takes place
as follows:
NaBH 4
8OH
NaBO2
6H 2 O e
E
1.24 V
(6.24)
The electrolyte in the DBFC is an alkali through which OH− ions migrate to the anode
from the cathode where they are produced by reduction of oxygen as in the traditional
AFC, i.e.,
O2
2H 2 O 4 e
4OH
0.4 V
E
(6.25)
The combination of reactions (6.24) and (6.25) yields a theoretical cell voltage of 1.64 V,
which is higher than any of the other DLFCs discussed in this chapter.
The sodium borate (NaBO2) produced by the direct anode reaction (6.24) is environmentally friendly and can be recycled back to sodium borohydride. The high cell
potential of the DBFC gives rise to a theoretical specific energy of 9.3 kWh kg−1, which
is greater than that of methanol (5.54 kWh kg−1). The specific energy will be lower when
using borohydride solutions despite the fact that these can be prepared up to a level of
30 wt.% in concentrated alkaline aqueous solutions (>6 M).
Unfortunately, the high anode potential predicted by reaction (6.24) is rarely achieved,
because on most metals borohydride spontaneously hydrolyses to generate a hydroxyl
borohydride intermediate and then hydrogen according to
BH 4
H2 O
BH3 OH
BH3 OH
H2 O e
H2
BO2
e
(6.26)
3H2
(6.27)
The presence of atomic hydrogen on the DBFC anode gives this electrode a mixed
potential as a consequence of reaction (6.24) and the competing reaction:
H2
2OH
2H 2 O 2e
E
0.828 V
(6.28)
The observed anode potential is therefore between −1.24 and −0.828 V.
Note that reactions (6.26) and (6.27) form the basis of the IBFC. They can be carried
out in a reactor that is fed with borohydride solution and are favoured by neutral or acid
conditions and promoted by catalysts.
An alternate means of generating OH− ions at the cathode of the DBFC, or indeed any
AFC, is by the reduction of hydrogen peroxide; the method has been proposed for
application in underwater vehicles and anaerobic systems. Research has demonstrated
that alkali systems employing hydrogen peroxide produce cell voltages that are higher
than those obtained when using air‐breathing cathodes.
179
180
Fuel Cell Systems Explained
6.6.1
Anode Catalysts
The electrochemical oxidation of borohydride was considered for fuel‐cell applications
in the 1960s using porous nickel and palladium anodes. Direct oxidation requires
selective anode catalysts with high activity for reaction (6.24) and low activity for the
hydrolysis reaction (6.23). Recent electrode materials that have been investigated
include N2B, Pd–Ni, Au, colloidal Au, Au alloys with Pt and Pd, MnO2, mischmetal,14
AB5‐type hydrogen storage alloys (see Chapter 11), Raney Ni, Cu, colloidal Osmium
and Osmium alloys. Only gold appears to achieve the transfer of eight electrons that
is predicted by reaction (6.24) and leads to the highest cell voltage. Nickel, while still
active, enables the transfer of just four electrons, i.e., about half of the theoretical
energy value from the fuel can be obtained with nickel anodes.
As gold is not able to absorb hydrogen, it is proposed that the oxidation reaction on this
metal proceeds via a mechanism in which the first step is the creation of borohydride
radicals BH4• by the extraction of electrons, i.e.,
BH 4
e
BH 4 •
(6.29)
The second step involves the oxidation of the radical to BH3− and water, followed by the
formation of diborane, B2H6, which undergoes further electron transfers.
On other metals such as Ni, Pt or Pd, the borohydride radical is dissociated on the
surface according to the following reaction where M represents the catalyst metal:
M BH 4 •
M BH3
M H
(6.30)
The adsorbed BH3− is then oxidized by surface and electron-transfer reactions. Most
metals do not achieve the transfer of eight electrons, which indicates partial oxidation
to intermediate products, or high rates of reaction (6.23).
A number of hydrogen storage alloys have been proposed as anode catalysts for the
borohydride fuel cell, e.g., ZrCr0.8Ni1.2, MmNi3.55Al0.3Mn0.4Co0.79 (Mm = mischmetal).
Hydrogen is generated by the borohydride, which is then stored in the lattice of the
alloy. Such electrodes have exhibited moderately high current densities (up to
300 mA cm−1) at 0.7 V but with low efficiency (i.e., only four electrons are transferred).
6.6.2
Challenges
Both cation‐ and anion‐permeable membranes have been tested for use in DBFCs
containing NaOH electrolyte. Each type of membrane leads to different chemistries
within the cell. Cation permeable membranes (i.e., permeable to Na+ ions) lead to a
chemical imbalance — the oxidation of 1 mol of NaBH4 transfers 8 mol of Na+ ions
across the membrane and thereby increases the concentration of NaOH in the cathode
region with concomitant decrease in the anode. Given the latter situation, extended
operation of the cell could raise problems because BH4− ions are stable only in solutions
with strong alkali concentration. With a cation membrane, it is therefore necessary to
14 Mm denotes mischmetal—an alloy of cerium, lanthanum, neodymium and other rare earth metals.
Direct Liquid Fuel Cells
(a)
Load
+
–
OH– OH–
6H2O
BO2–
8e–
12OH–
8e–
8Na+
Anode
8OH–
BH4–
2O2
4H2O
Cathode
OH– OH–
Cation-permeable membrane
(b)
Load
+
–
OH– OH–
8e–
Anode
6H2O
BO2–
8OH–
BH4–
12OH–
8OH–
OH– OH–
2O2
4H2O
8e–
Cathode
Anion-permeable membrane
Figure 6.8 Principles of operation of a DBFC with (a) a cation‐conducting membrane and (b) an anion‐
conducting membrane. Drawn to emphasize the chemical balance of the reactions at the electrodes.
(Source: From Ponce de Leon, C and Walsh, FC, 2009, Sodium borohydride fuel cells, in Garche, J, Dyer,
CK, Moseley, PT, Ogumi, Z, Rand, DAJ and Scrosati, B (eds.), Encyclopedia of Electrochemical Power
Sources, pp. 192–205, Elsevier, Amsterdam. Reproduced with the permission of Elsevier.)
recycle NaOH from the cathode to the anode. A diagrammatic representation of the
operation is given in Figure 6.8a.15
By contrast, an anion permeable membrane transfers 8 mol of OH− from cathode to
anode across the membrane for each mole of borohydride that is oxidized. The chemistry
using this membrane is in balance, and to maintain power production only borohydride
needs to be supplied to the anode, as shown in Figure 6.8b.
The cation permeable membranes that have been the most widely investigated for
DBFCs are the various Nafion materials used in PEMFCs. These are stable in alkaline
solution. By contrast, at present, there is no commercially available anionic membrane
material that can survive the high alkalinity necessary to maintain NaBH4 in solution
without it being hydrolyzed to hydrogen.
It is, of course, possible to operate a DBFC with a liquid alkaline electrolyte and no
ion‐conducting membrane. Sodium or potassium hydroxide solutions with concentrations
15 Ponce de Leon, C and Walsh, FC, 2009, Sodium borohydride fuel cells, in Garche, J, Dyer, CK, Moseley,
PT, Ogumi, Z, Rand, DAJ and Scrosati, B (eds.), Encyclopedia of Electrochemical Power Sources, pp. 192–205,
Elsevier, Amsterdam.
181
182
Fuel Cell Systems Explained
between 10 and 40 wt.% work well (in which borohydride solutions of 10–30 wt.% are
used as fuel). With these electrolytes, however, crossover becomes a problem but is less
of a concern with cells containing anionic membranes. In anionic membrane cells, BH4−
ions are prevented from reaching the cathode by the membrane. Although there is
tendency, particularly at low currents, for BH4− ions to flow towards the cathode, flow
of OH− ions in the opposite direction keeps the cell charges in balance.
6.7
Application of Direct Liquid Fuel Cells
Currently, the DMFC is the only DLFC that can be said to have reached a stage of
sustained commercialization. Several educational kits and specialized systems of other
direct liquid systems, such as the DEFC marketed by Horizon Fuel Cells, can also be
purchased, but these do not constitute a large market segment.
State‐of‐the‐art DMFCs can achieve power densities of up to 60 mW cm−2. This is
considerably lower than the performance of hydrogen fuel cells and constrains the area
of application to duties where the power density can be low, but the energy density must
be high. To put it another way, DMFCs are suited to services where the average power
is only a few watts, but that power must be provided for a very long time — typically, for
several days. Example applications include mobile phones, laptops, remote monitoring
and sensing equipment, and mobile homes. In the case of consumer electronics, increasing
computing power is placing heavy demands on batteries that are driving improvements
in lithium‐ion technology. The best state‐of‐the‐art lithium batteries pack around
0.6 Wh mL−1 of energy. By comparing this performance with the high energy densities
of fuels given in Table 6.1, it is not difficult to see why the DLFCs are also of great
interest. The energy density of the liquid alone is significantly higher than that of the
battery. Even when the conversion efficiency of the fuel cell is taken into account, the energy
densities of fuel‐cell systems are still higher than those of batteries. Moreover, this
ignores a distinguishing feature of the fuel cell, namely, that it will continue to produce
power as long as fuel is supplied. The size of the fuel cell is therefore determined by the
maximum power in watts that it needs to supply16 for the particular application. By
contrast, the size of a rechargeable battery is governed by the watt‐hours that it needs
to provide. It may take several hours to recharge a lithium battery that has been
discharged to, say, a 10% state-of-charge. There is no issue with the recharging of a fuel
cell. A canister of methanol that may power a laptop through a DMFC system can be
replaced in less than a minute once it becomes drained.
Many DMFC stacks that employ Nafion membranes have been demonstrated for
portable applications by organizations such as Motorola Labs, Energy Related Devices,
Samsung Advanced Institute of Technology, Los Alamos National Laboratory and
the Jet Propulsion Laboratory, and by various research groups in universities. SFC
Energy AG in Germany has the longest track record in the commercialization of
16 Flow batteries such as the vanadium redox battery or the zinc bromine battery, introduced in Chapter 1,
are also sized according to kW rather than kWh — another reason they are often categorized as ‘fuel cells’
rather than as ‘batteries’.
Direct Liquid Fuel Cells
(a)
(b)
Figure 6.9 (a) EFOY Comfort fuel cell installed in a mobile home with a container of methanol on the
left and (b) range of EFOY Pro series (800–2400 W). (Source: Reproduced with permission of Elsevier.)
DMFC systems. The EFOY Comfort series of DMFCs produced by this company covers
systems from 40 to 85 W nominal output that consume methanol at a rate of 0.9 L
kWh−1. One of these systems located in a mobile home is shown in Figure 6.9a, together
with a range of larger EFOY Pro series DMFCs in Figure 6.9b. Oorja Protonics (USA) is
supplying 1‐kW DMFC systems for stationary power and materials handling
applications (forklift trucks), and a 5‐kW system has been developed under a European
‘Dreamcar’ project17 for application in vehicle auxiliary power units, as shown in
Figure 6.10.
17 Liu, H and Zhang, J (eds.), 2009, Electrocatalysis of Direct Methanol Fuel Cells, WILEY‐VCH Verlag
GmbH & Co. KGaA, Weinheim.
183
184
Fuel Cell Systems Explained
Figure 6.10 5‐kW DMFC stack developed within the framework of the European ‘Dreamcar’ project.
(Source: Reproduced from Arico, AS, Baglio, V and Antonucci, V, 2009, Direct methanol fuel cells:
history, status and perspectives, in Liu, H and Zhang, J (eds.), Electrocatalysis of Direct Methanol Fuel
Cells, WILEY‐VCH Verlag GmbH & Co. KGaA, Weinheim. Reproduced with permission of Wiley‐VCH.)
Further Reading
Adamson, K‐A and Pearson, P, 2000, Hydrogen and methanol; a comparison of safety,
economics, efficiencies, and emissions, Journal of Power Sources, vol. 86, pp. 548–555.
Arico, AS, Baglio, V and Antonucci, V, 2009, Direct methanol fuel cells: history, status and
perspectives, in Liu, H and Zhang, J (eds.), Electrocatalysis of Direct Methanol Fuel
Cells, WILEY‐VCH Verlag GmbH & Co. KGaA, Weinheim.
Badwal, S, Giddey, S, Kulkarni, A and Jyoti, G, 2015, Direct ethanol fuel cells for transport
and stationary applications – a comprehensive review, Applied Energy, vol. 145,
pp. 80–103.
Choi, WC, Kim, JD and Woo, SI, 2001, Modification of proton conducting membrane for
reducing fuel crossover in a direct methanol fuel cell, Journal of Power Sources, vol. 96,
pp. 411–414.
Dohle, H, Divisek, J and Jung, R, 2000, Process engineering of the direct methanol fuel cell,
Journal of Power Sources, vol. 86, pp. 469–477.
Dohle, H, Schmitz, H, Bewer, T, Mergel, J and Stolten, D, 2002, Development of a
compact 500W class direct methanol fuel cell stack, Journal of Power Sources, vol. 106,
pp. 313–322.
Dyer, CK, 2002, Fuel cells for portable applications, Journal of Power Sources, vol. 106,
pp. 31–34.
Hamnett, A 1997 Mechanism and electrocatalysis in the direct methanol fuel cell,
Catalysis Today, vol. 38, pp. 445–457.
Direct Liquid Fuel Cells
Jarvis, LP, Terrill, BA and Cygan, PJ, 1999, Fuel cell/electrochemical capacitor hybrid for
intermittent high power applications, Journal of Power Sources, vol. 79, pp. 60–63.
Jung, DH, Cho, S, Peck, DH, Shin, D and Kim, JJ, 2002, Performance evaluation of a
Nafion/silicon oxide hybrid membrane for direct methanol fuel cell, Journal of Power
Sources, vol. 106, pp. 173–177.
185
187
7
Phosphoric Acid Fuel Cells
7.1
High‐Temperature Fuel‐Cell Systems
In Chapter 2, it was noted that the open‐circuit voltage for a hydrogen fuel cell decreases
at higher temperatures. Indeed, above about 800°C, the theoretical maximum efficiency
of a fuel cell is actually less than that of a heat engine. On this basis, one may question
why fuel cells should be operated at higher temperatures? The reason is that, in many
cases, high temperatures bring the following benefits that outweigh the disadvantages:
●
●
●
●
Electrochemical reactions proceed more rapidly at higher temperatures, and thus
voltage losses due to electrokinetic (‘activation’) effects are lower. Consequently,
precious metal catalysts are often not required.
The exhaust gases from the fuel‐cell stacks are sufficiently hot to facilitate the generation
of hydrogen from other fuels that are readily available, e.g., natural gas.
The exhaust gases are at a high temperature and therefore a valuable source of heat
for buildings, processes and facilities near the fuel‐cell installation. In other words,
these types of fuel cell make excellent ‘combined heat and power’ (CHP) systems.
Heat extracted from exhaust gases and cooling fluids can be employed to drive turbines and generators to produce more electricity. When a turbine uses waste heat
from a generator, such as a fuel cell, the scheme is known as a ‘bottoming cycle’.1
A combination of a fuel cell and heat engine allows the complementary characteristics
of each to be exploited to great advantage, so that electricity can be generated with a
higher level of efficiency.
The phosphoric acid fuel cell (PAFC) is the most developed of the common competing
types of technology that operate at temperatures above about 200°C. Many 200‐kW
PAFC CHP systems are installed throughout the world at hospitals, military bases,
leisure centres, offices, factories and even prisons. Their performance and behaviour
are well understood. The moderate operating temperature of the PAFC requires the use
of noble metal catalysts, and, as with the PEMFC, these will be poisoned by any carbon
monoxide (CO) that may be in the fuel gas. A somewhat complex fuel‐processing
system is required to achieve acceptably low levels of CO.
1 Conversely in a ‘topping cycle’, electricity is produced primarily from a steam turbine. The exhaust steam
from the turbine is condensed, and the heat released is utilized in external applications such as district
heating or water desalination.
Fuel Cell Systems Explained, Third Edition. Andrew L. Dicks and David A. J. Rand.
© 2018 John Wiley & Sons Ltd. Published 2018 by John Wiley & Sons Ltd.
188
Fuel Cell Systems Explained
Some system design issues are common to all high‐temperature fuel cells and are worth
addressing before examining PAFC systems in detail. These issues relate principally to
the fate of the heat generated by the fuel cells, namely, whether the heat is used to
reform fuels, drive engines or facilitate practical applications. Thus in evaluating the
utility of PAFC, molten carbonate fuel cell (MCFC) and solid oxide fuel cell (SOFC)
stacks, each of the three technologies should not be considered in isolation but rather
as an integral component of a complete system that generates both heat and power.
The common features for the high‐temperature fuel cells are as follows:
●
●
●
●
A PAFC, MCFC or SOFC will nearly always use a fuel that will need refining or
processing. A detailed review of fuel processing is given in Chapter 10, but the basics
of how this operation is integrated into the fuel‐cell system and subsequently impinges
on overall performance is explained in Section 7.2.1.
The fuel will invariably be a mixture of hydrogen, carbon oxides and other gases.
During passage of the fuel gas through the stack, hydrogen will be consumed, and the
resulting reduction in its concentration in the mixture will lower the local current
density. ‘Fuel utilization’ is an important operating parameter and is discussed in
Section 7.2.2.
The high‐temperature exhaust gases carry large amounts of heat energy that can
be employed in a bottoming cycle with a turbine or other heat engines. How this
combination of fuel cell and heat engine can lead to very high levels of efficiency is
considered in Section 7.2.3.
Heat from the exhaust gases can also serve to preheat fuel and oxidant with the aid
of suitable heat-exchangers. The best use of heat within high‐temperature fuel‐cell
systems is an important aspect of system design and is often referred to as ‘process
integration’ by chemical engineers. To achieve high electrical and thermal efficiencies,
systems need to be designed to minimize exergy loss, and designers may introduce
‘pinch technology’ to achieve the best outcome for process integration. Such system
heat management aspects are covered in Section 7.2.3.
7.2
System Design
7.2.1
Fuel Processing
As this topic is described in detail in Chapter 10, it is sufficient at this stage to say that
the production of hydrogen from a hydrocarbon usually involves the process of ‘steam
reforming’. The procedure should not be confused with the reforming of hydrocarbons
as practised in the petroleum industry. In the case of methane, the steam reforming
reaction (often referred to as SMR) may be written as:
CH 4
H2 O
3H2
CO
(7.1)
A general expression to include other hydrocarbons, represented by C xHy, can be
written as:
CxHy
xH 2 O
x
y
H2
2
xCO
(7.2)
Phosphoric Acid Fuel Cells
In most cases, and certainly with natural gas, the SMR is ‘endothermic’. That is, heat
needs to be supplied to drive the reaction forward to produce hydrogen. Again, for
virtually all fuels, the reforming has to be conducted at relatively high temperatures,
usually well above about 500°C. With medium‐ and high‐temperature fuel cells, heat
required by the reforming reactions can be provided, at least in part, from the fuel cell
itself, i.e., from the exhaust gases. In the case of the PAFC, the heat at around 200°C has
to be supplemented by burning fresh fuel gas. This requirement lowers the efficiency
of the overall system so that, for the PAFC, the upper limit falls to 40–45% (LHV). By
comparison, heat carried by the exhaust gases from both the MCFC and the SOFC is
available at much higher temperatures. If all of the exhaust heat from the MCFC or
SOFC stack is used to promote the SMR (especially when the process is performed
inside the stack), the outcome is high system efficiency. Typically >50% (LHV) efficiency
is achievable for MCFC or SOFC systems.
For the PAFC, as with the PEMFC, the gas mixture produced by steam reforming
must be further processed to reduce the concentration of CO in the mixture. The
‘water-gas shift’ reaction (usually abbreviated as the ‘shift reaction’), whereby CO is
converted to carbon dioxide, is employed, namely:
CO H2 O
CO2
H2
(7.3)
This process is generally carried out in two stages (see Section 10.4.9, Chapter 10) in
reactors that operate at different temperatures— to achieve levels of CO that are
sufficiently low to be acceptable for PAFC stacks.
A further complication is that fuels such as natural gas nearly always contain small
amounts of sulfur or sulfur‐containing compounds. Sulfur is a well‐known catalyst
poison, i.e., it will absorb preferentially on the catalyst metal and reduce the activity for
both steam reforming and shift reaction. In a similar manner, sulfur will also deactivate
the electrode catalysts of all types of fuel cell. Consequently, it is essential that this
impurity is removed from the fuel gas before it is fed to the reformer or stack.
Desulfurization is well established industrially and is featured in many hydrocarbon
processes, not just for fuel cells; the process is discussed further in Section 10.4.2,
Chapter 10.
7.2.2
Fuel Utilization
The issue of fuel utilization arises whenever the hydrogen for a fuel cell is supplied as
one component of a reactant gas or becomes a component of the gas mixture due to
internal reforming. Consider a purified fuel gas for a PAFC containing hydrogen, carbon dioxide and water vapour. As this gas mixture flows through a cell, the hydrogen is
consumed electrochemically, and CO2 and H2O simply pass through without reacting.
The result is that the partial pressure of hydrogen falls as the fuel gas travels from cell
inlet to outlet. A similar effect is observed with oxygen in the air on the cathode side of
the cell.
The effects of pressure and gas concentration on the open‐circuit voltage of a fuel cell
have been examined in Section 2.5, Chapter 2. The Nernst relationship, introduced
as equation (2.36), relates the open‐circuit voltage, Vr, and the partial pressures of
hydrogen, oxygen and steam as follows:
189
190
Fuel Cell Systems Explained
Vr
Vr
PH
PO
2
2
.
RT
P
ln P
2F
PH O
1
2
(2.36)
2
P
If only the partial pressure of hydrogen is considered and the pressure changes from
Pin to Pout, then the change in cell voltage is expressed by:
V
P
RT
ln out
2F
Pin
(7.4)
Given that the partial pressure of hydrogen in the fuel gas is falling due to the reaction
taking place within the cell, Pout is always less than Pin, and thus ΔV will always be negative.
The open‐circuit cell voltage, and therefore the voltage under load, could be expected
to fall on moving from inlet to outlet. This clearly cannot be the case since bipolar plates
are good electronic conductors, and therefore the voltage difference between the two
electrodes of a fuel cell must be the same over the whole area of the cell. The local cell
voltage under load measured at the fuel inlet must be the same as that measured at the
fuel outlet. For this situation to occur, the local current density must be lower at the
outlet of the cell than at the inlet to accommodate the fact that less hydrogen is available
to react at the outlet of the cell compared with the inlet.
The above situation is assumed to hold especially for the SOFC and the MCFC in
which the activation overpotentials at each electrode are relatively small and the internal
ohmic losses are taken to be uniform throughout the cell. Recent research has found
that this is not necessarily the case for PEMFCs. Careful in situ measurements have
shown variations in both current density and local impedance for these fuel cells according to position in the cell. Example data for the two parameters in a cross‐flow PEMFC
are given in Figure 7.1. Current density is highest towards the corner of fuel and oxidant
inlets and also high towards the corner where both fuel and oxidant exit the cell—
confirmed as segments 17 and 33 in Figure 7.1a. By contrast, the AC impedance spectra
for these two segments given in Figure 7.1b are clearly very different, showing that there
is difference in impedance depending on cell position. Indeed, in this example, the
segment in the centre of the cell (25) exhibits a similar impedance spectrum to that at
the cell outlet (33).
Equation (7.4) shows that ΔV is also dependent on temperature, which means that the
expected open‐circuit voltage drop, and hence the reduction in current density as a
result of the falling partial pressure of hydrogen through the anode, will be greater for
fuel cells operating at higher temperatures.
At the cathode of the fuel cell, the partial pressure of oxygen in the air will also reduce
as it passes through the cell. This is less of a problem in practical terms since the cell
voltage is dependent on the square root of the partial pressure of oxygen, as indicated
by equation (2.36). The influence of fuel and oxygen utilization on the open‐circuit
voltage, Vr, is illustrated in Figure 7.2. The uppermost dashed line shows the voltage
of a typical hydrogen fuel cell, which is operating at 100 kPa and supplied with pure
hydrogen and oxygen. The lower dashed line is for a cell using air at the cathode and a
Phosphoric Acid Fuel Cells
(a)
λair = 3
0.050
Seg.33
H2 inlet
Seg.25
0.150
Seg.17
Acm–2
0.100
0.200
0.250
0.300
Air inlet
(b)
0.10
λair = 3
Seg. 17
Seg. 25
Seg. 33
0.05
Im(Z)/Ωcm2
0.00
–0.05
–0.10
–0.15
–0.20
0.1
0.2
0.3
0.4
0.5
0.6
0.7
Re(Z)/Ωcm2
Figure 7.1 (a) Current density distribution and (b) electrochemical impedance spectroscopy Nyquist
plot at selected cell segments for a PEMFC at 200°C; cathode stoichiometry of λ = 3. (Source: Bergmann,
A, Kurz, T, Gerteisen, D and Hebling, C, 2010, Spatially resolved impedance spectroscopy in PEM fuel
cells up to 200°C, in: Stolten, D. and Grube, T. (eds.), 18th World Hydrogen Energy Conference, WHEC
2010, Parallel Sessions Book 1: Fuel Cell Basics / Fuel Infrastructures, Proceedings of the WHEC, May
16–21.Reproduced with permission of Forschungszentrum J¸lich GmbH, Zentralbibliothek, Verlag.)
mixture of four parts hydrogen to one part carbon dioxide at the anode to simulate the
gas mixture that would be obtained from reformed methane. The upper and lower solid
lines are plots of the ‘open‐circuit exit voltage’ for 80 and 90% fuel utilization, respectively, with 50% air consumption in both cases. The data are for a situation in which
both the air and fuel are flowing in the same direction (co‐flow). Under such conditions,
the drop in the open‐circuit voltage is significant for a co‐flow configuration and, as
expected, increases with both temperature and fuel utilization.
Sometimes, the current density distribution through a high‐temperature cell can be
made more uniform by feeding the air and fuel through the cell in opposite directions,
i.e., a counter‐flow operation. With this arrangement, the fuel outlet region of the cell
has the highest partial pressure of oxygen. It should be noted, however, that the particular
191
Fuel Cell Systems Explained
1.2
Pure reactants
1.1
Open-circuit voltage / V
192
Entry voltage for air and
reformed methane fuel
1.0
0.9
Voltage at exit, 80%
fuel utilization
0.8
Open-circuit voltages at cell exit
Voltage at exit, 90%
fuel utilization
0.7
0.6
0
200
400
600
800
1000
Temperature/°C
Figure 7.2 Open‐circuit voltage of a hydrogen fuel cell under different conditions. The two curves for
the voltage at the exit show how the voltage depends on fuel utilization and temperature. In both
cases, the oxygen utilization is 50%.
configuration adopted for a fuel cell is also dependent on how the fuel and air flows
influence the temperature distribution within the stack. This, in turn, is influenced by
the method adopted for stack cooling.
It should be remembered that steam is produced at the anode in both an MCFC and
an SOFC rather than at the cathode as in a PEMFC and a PAFC. In other words, in an
MCFC or SOFC, the hydrogen in the fuel is essentially replaced, as it is consumed, by
steam. Accordingly, if the partial pressure of the hydrogen decreases as it passes through
the fuel‐cell anode compartment, the partial pressure of steam will increase, and, as
previously indicated by equation (2.36), the outcome will be a fall in the open‐circuit
voltage. Unfortunately, the effect of this behaviour is difficult to model because, for
example, some of the steam may be employed in internal fuel reforming. The situation
is therefore liable, in practice, to be worse than Figure 7.2 would indicate.
It can be concluded that, in the case of a reformed fuel containing carbon dioxide or
when internal reforming is applied, it is impossible to consume all the hydrogen in the
fuel‐cell stack. Some of the hydrogen must therefore pass straight through the cell
unconverted, to be used later to provide energy to process the fuel or to be burnt to
increase the heat energy available for further operations, as discussed in Section 7.2.3.
In the early days of fuel‐cell development, optimum values of both fuel and air utilization
in PAFCs, MCFCs and SOFCs were determined experimentally. The task has been
made easier in recent years by the advent of computer models that can simulate entire
fuel‐cell systems.
7.2.3
Heat‐Exchangers
Not surprisingly, there are challenges in the way in which the various components of a
fuel‐cell system are integrated. The situation applies to all types of fuel cell, but is
Phosphoric Acid Fuel Cells
particularly notable for PAFC, MCFC and SOFC systems where several of the balance‐
of‐plant items operate at high temperatures. Examples of such items are the desulfurizer,
reformer reactor, shift reactors, heat-exchangers, recycle compressors and ejectors. In
some of these components, heat may be generated or consumed. The challenge for the
system designer is to arrange the various components in a manner that minimizes heat
losses to the external environment and at the same time ensures that heat is utilized
within the system in the best possible way (i.e., by avoiding unnecessary losses).
7.2.3.1
Designs
In any fuel‐cell system, heat is required by several process streams, e.g., for preheating
the fuel fed to the reformer, for running the reformer itself and for raising and superheating steam. There are also areas that need to be cooled, e.g., the fuel‐cell stack and,
in the case of a PEMFC, also the outlet of the shift reactor(s). Heat transfer from one
process stream to another is carried out by means of a heat-exchanger. The gas (or liquid)
to be heated passes through pipework that is heated by the gas (or liquid) to be cooled.
A commonly used symbol for a heat-exchanger is shown in Figure 7.3. When the exit
fluids from a process are employed to heat incoming fluids, the heat-exchanger is often
called a ‘recuperator’.
There are several types of heat-exchanger that include the shell‐and‐tube, plate‐fin
and printed‐circuit designs. The selection for any particular application will be governed
by the temperature range of operation, the fluids involved (e.g., liquid or gas phase), the
fluid throughput and the cost. The materials of construction, the method of fabrication
and the heat-transfer area required for the application determine the cost of the
exchanger.
7.2.3.2
Exergy Analysis
Exergy is the maximum amount of work that can be done by a system as it approaches
thermodynamic equilibrium with its surroundings by a sequence of reversible processes.
Consequently, the exergy of a system can be considered as a measure of its ‘distance’
from equilibrium with the surroundings. When the system and its surroundings are in
equilibrium, the exergy of the system is
zero. Therefore, thermal exergy is simply
‘available heat’. Potential energy and
Fluid losing
heat
kinetic energy, as classically defined, are
also forms of exergy as is the Gibbs free
energy of combustion of a fuel (with the
sign changed). Energy is conserved in all
processes (first law of thermodynamics)
whereas exergy is conserved in processes
Fluid gaining
that are reversible. Real processes are of
heat
course irreversible, so that exergy is always
partly consumed to give enthalpy.
In a power conversion device such as a
fuel‐cell system where a chemical reaction
is continually taking place, the state of the
system cannot be defined merely from the
Figure 7.3 Common symbol for a heattemperature, the volume and the pressure.
exchanger. The fluid to be heated passes through
In recognition of this, Gibbs defined a
the zigzag element.
193
194
Fuel Cell Systems Explained
property, μ, known as the chemical potential of a substance or system. When there is an
energy change in a system that also involves a chemical reaction, the change in Gibbs
free energy can be represented formally by:
G V P S T
i
(7.5)
ni
where V, S and ni denote extensive parameters of the system (volume, entropy and number of moles of different chemical components, respectively); P, T and μi are intensive
parameters of the environment (pressure, temperature and chemical potential of the
components). It can be shown2 that the change in exergy (ΔB) of a system in going from
an initial state to a reference state (subscript o) is given by
B S T To
V P Po
ni
i
o
(7.6)
Clearly the higher the temperature, the greater is the exergy of the system.
Consider, for instance, the case where a PEMFC and an SOFC system have the same
power output and efficiency. The heat, i.e., the enthalpy content, of the exhaust streams
from both systems will be the same. The heat that is liberated in a PEMFC is at a
temperature of around 80°C and thereby is of limited value both within the system and
for external applications. For the latter, it may be applied to space heating in buildings
or possibly integration with an absorption cooling system to provide air cooling. In the
design of a PEMFC system, care should be taken to ensure that heat is used efficiently
so that there is maximum available exhaust heat. By contrast, the heat produced by the
SOFC will be at a much higher temperature and therefore will have a higher exergy and,
consequently, be more valuable for further utilization than that from a PEMFC. For
example, the exhaust heat from an SOFC could enable the powering of a steam turbine
in a bottoming cycle.
All fuel‐cell systems should therefore be configured in such a way that exergy loss is
minimized. This is especially important for PEMFC and PAFC systems, which operate
at moderate temperatures, where any heat utilized inefficiently within the system
will have a more deleterious effect on the amount of exhaust heat that is available
externally.
7.2.3.3
Pinch Analysis
Pinch analysis, or pinch technology, is a methodology that can be applied to fuel‐cell
systems for deciding the optimum arrangement of heat-exchangers and other units so
as to minimize loss of exergy. It was originally designed by chemical engineers as a tool
for defining energy‐saving options, particularly in heat-exchanger networks, but has
since been applied in the development of fuel‐cell systems. The concept is fairly straightforward but for complex systems, sophisticated computer models are required. The
procedure for pinch technology is broadly as follows.
In any fuel‐cell plant, there will be process streams that require heating (cold
streams) and cooling (hot streams), irrespective of where heat-exchangers are located.
The first stage in system design is therefore to establish the basic chemical‐processing
requirements and to produce a configuration that shows and defines all of the cold
2 Dincer, I and Cengel, YA, 2001, Energy, entropy and exergy concepts and their roles in thermal
engineering, Entropy, vol. 3, pp. 116–149.
Phosphoric Acid Fuel Cells
700
600
Cathode
feed
Cathode effluent
Cathode
feed
Temperature/°C
500
400
Combustion
air preheat
Process
+ motive ss
Flue gas
300
Process ss
200
Recycle gas
preheat
Feed + fuel gas
100
preheat
Bfw
preheat
Wet flue gas
Dearation preheat
0
0
2
4
6
8
10
12
14
16
Enthalpy/kW x 100
18
20
22
24
26
Figure 7.4 Hot and cold heating plots for a conceptual 3.25‐MW MCFC system with high‐pressure
steam generation. ss, steam superheat; Bfw, boiler feed water.
and hot streams. Calculation of the heat and mass balances enables the engineer to
determine the enthalpy content of each stream. Heating and cooling curves can then
be produced from knowledge of the required temperatures of each stream; examples
for an MCFC system are given in Figure 7.4. The individual cooling and heating
curves are then summed together to make two composite plots—one shows the total
heating required by all of the streams that need heating and the other shows the total
cooling required by the streams that require cooling. The composite plots obtained
by summing the curves of Figure 7.4 are given in Figure 7.5. The composite plots are
brought together by sliding along the x‐axis and where they ‘pinch’ together with a
minimum temperature difference of, for example, 50°C; the temperature is noted.
This so‐called pinch point defines the target for optimum process design, since in a
real system heat cannot be transferred from above or below this pinch temperature.
Once the pinch point is known, heat-exchangers can be positioned in such a way that
maximum transfer of heat is achieved between units that need heating and those
that require cooling. In some fuel‐cell systems, a pinch temperature is not found, in
which case the problem becomes one of defining an upper limit of temperature for
the system. Either way, pinch technology provides an excellent method for system
optimization. Many computer models are available for calculating the heat and material
flows around the system and for calculating the pinch‐point temperature.3 Once
such an analysis has been undertaken, the required heat-exchangers and reactors can
be designed.
3 Aspen Technology Inc., for example, produces a suite of software packages including Aspen Plus® and
AspenTech Pinch™ that are widely used for process and system design.
195
Fuel Cell Systems Explained
800
Cold stream
Hot stream
700
600
Temperature/°C
196
500
Pinch point
400
300
200
100
0
0
2
4
6
8 10 12 14 16 18 20 22 24 26 28 30 32
Enthalpy/kW x 100
Figure 7.5 Composite curves derived from data presented in Figure 7.4.
Of course, other considerations need to be taken into account when designing a fuel‐
cell plant, for example, the choice of materials for the balance‐of‐plant components and
the mechanical layout of the system. Configurations are drawn up initially as a process
flow diagram (PFD) that shows the logical arrangement of the fuel‐cell stack and the
associated balance‐of‐plant components. Many other techniques that are available to
the system designer are outside the scope of this book. Nonetheless, it is hoped that this
consideration of common elements of high‐temperature fuel‐cell systems will be a good
starting point.
7.3
Principles of Operation
The PAFC works in a similar fashion to the PEMFC, as described in Chapter 4. A
proton‐conducting electrolyte is employed, and the reactions that occur on the anode
and cathode are those given in Figure 1.3, Chapter 1. The electrochemical reactions take
place on highly dispersed electrocatalyst particles that are supported on carbon black. As
with the PEMFC, the PAFC uses platinum (Pt) alloys as the catalyst at both electrodes.
The electrolyte is an inorganic acid, concentrated phosphoric acid (100 wt.%).
7.3.1
Electrolyte
Phosphoric acid (H3PO4) is the only common inorganic acid that, above 150°C, has
satisfactory thermal, chemical and electrochemical stability and sufficiently low volatility to be considered as an electrolyte for fuel cells. Most importantly, phosphoric
acid is tolerant to carbon dioxide in the fuel and oxidant, unlike the electrolyte solution
in the alkaline fuel cell. The acid was therefore chosen by United Technologies
Phosphoric Acid Fuel Cells
(a US company, later becoming the spin‐off ONSI Corporation) in the 1970s as the
preferred electrolyte for fuel‐cell power plants in terrestrial applications.
Phosphoric acid is a colourless, viscous, hygroscopic liquid. It is contained in the
PAFC by capillary action (it has a contact angle >90°) within the pores of a matrix made
of silicon carbide (SiC) particles of about 1 µm that are held together with a small
amount of PTFE. Pure 100% phosphoric acid, which has been used in fuel cells since the
early 1980s, has a freezing point of 42°C. Therefore, to avoid the development of stresses
due to freezing and rethawing, PAFC stacks are usually maintained above this temperature
once they have been commissioned. Although the vapour pressure is low, some acid is
lost during normal fuel‐cell duty over long periods at high temperature; the amount
depends on the operating conditions, especially gas flow velocities and current density.
It is therefore necessary either to replenish electrolyte during service or to ensure that
at the start of operation, there is sufficient reserve of acid in the cell to sustain the
projected lifetime. The SiC matrix is thin enough (0.1–0.2 mm) to keep ohmic losses at
a reasonably low level (i.e., to give high cell voltages) while having adequate mechanical
strength and the ability to prevent crossover of reactant gases from one side of the
cell to the other. This latter property is a challenge for all fuel cells with liquid‐based
electrolytes. Under some conditions, the pressure difference between the anode and the
cathode can rise considerably, as determined by the design of the system.
Loss of phosphoric acid from the fuel cell can occur through volume change or
evaporation and transfer by electrochemical pumping. During operation, the volume
of phosphoric acid electrolyte expands and contracts according to temperature,
pressure, changes in load and the humidity of the reacting gases. A similar effect occurs
with the electrolyte in the MCFC, as will be discussed in Section 8.2, Chapter 8. To
replace electrolyte that may be lost through expansion or volume change, the porous,
carbon‐ribbed, flow‐field plates in the PAFC act as reservoirs for excess electrolyte.
The porosity and pore‐size distribution of these plates are deliberately chosen to
accommodate any volume changes of the electrolyte. Loss of electrolyte through
evaporation is minimized by keeping the stack operating temperature reasonably low,
but even at 200°C there is some escape of electrolyte through the air channels. In
practical PAFC stacks, evaporative loss is curtailed by ensuring that the cathode exit
gases pass through a cool condensation zone at the edge of the cell. With additional
cooling, the zone is maintained between 160 and 180°C, which is sufficiently low to
condense out most of the electrolyte vapour.
Electrochemical pumping is a phenomenon that occurs per se with fuel cells that
employ any liquid electrolyte or dissolved electrolyte. In the case of the PAFC, the
electrolyte dissociates into positively charged cations (H+) and negatively charged
anions (H2PO4−). During operation, the protons move from anode to cathode, whereas
the anions move in the other direction. Phosphate anions therefore build up at the
anode and can react with hydrogen to form phosphoric acid, thereby leading to an
accumulation of electrolyte at the cathode of each cell. Not surprisingly, a similar effect
can occur in the MCFC where pumping of alkali metal ions from the anode to cathode
causes a build‐up of carbonate electrolyte at the cathode of the fuel cell. Electrochemical
pumping can be minimized by the optimization of the porosity and pore‐size distribution
of the porous components, i.e., the ribbed plates and the electrodes in the case of
the PAFC and the MCFC, respectively. Degradation of the separator plate in the PAFC
can lead to migration of electrolyte from one cell to the next with consequent
197
198
Fuel Cell Systems Explained
catastrophic loss of stack voltage. Migration of acid can also occur through breakdown
of the manifold seals, again resulting in serious stack failure.
7.3.2
Electrodes and Catalysts
Like the PEMFC, the PAFC has gas‐diffusion electrodes, in which the catalyst is
platinum supported on carbon black. This catalyst has replaced the PTFE‐bonded
platinum black that was used in the first PAFC stacks that were built in the mid‐1960s.
In a modern PAFC, the catalyst layers contain 30–50 wt.% PTFE to act as a binder for
the creation of a porous structure. Meanwhile, the carbon catalyst support provides the
following functions similar to those fulfilled by the support in PEMFC catalysts:
●
●
●
To disperse the platinum to ensure good utilization.
To provide micropores in the electrode for maximum gas diffusion to the catalyst and
the electrode–electrolyte interface.
To increase the electrical conductivity of the catalyst layer.
The activity of the PAFC catalysts in both positive and negative electrodes depends on
the nature of the Pt, i.e., its crystallite size and specific surface area. In state‐of‐the‐art stacks,
the loadings are currently about 0.10 and about 0.50 mg Pt cm−2 in the anode and cathode,
respectively. The low loadings are, in part, the result of advances in nanotechnology—the
ability to prepare small crystallite sizes of around 2 nm in diameter with high specific
surface areas of up to 100 m2 g−1; see Figure 1.6, Chapter 1.
Each catalyst layer in the PAFC is usually bonded to a thin gas‐diffusion layer (GDL)
or substrate made of carbon paper. A typical GDL used in PAFCs has carbon fibres
of 10 mm in length that are embedded in a graphitic resin. The paper has an initial
porosity of about 90%, which is reduced to about 60% by impregnation with 40 wt.%
PTFE. The resulting wet‐proof carbon paper contains macropores of 3–50 µm diameter
(median pore diameter of about 12.5 µm), which can serve as a reservoir for phosphoric
acid, and micropores with a median pore diameter of about 3.4 nm to permit gas
permeability.
The composite structure of a carbon black + PTFE layer on a carbon paper substrate
forms a stable, three‐phase interface in the fuel cell, with electrolyte on the electrocatalyst side and the reactant gas environment on the other (carbon paper) side.
The choice of carbon for the catalyst layer is important, as is the method of dispersing
platinum, and much of the expertise in these two areas is proprietary to the manufacturers
of fuel cells. Through many decades of proof of operation in the field, the PAFC has
demonstrated good long‐term reliability. For instance, stack operating times can extend
well beyond 40 000 h before decay in electrode performance has reached an unacceptably
low level.
Phosphoric acid electrodes can be poisoned by carbon monoxide although the
tolerance is significantly greater than for PEMFC catalysts. Thus compared with the anode
catalyst of a PEMFC, which can accept up to only a few ppm of CO in the fuel gas, the
PAFC anode catalyst can tolerate typically up to about 2 mol.% at 200°C. In addition to
sulfur, which also poisons the catalyst, small amounts of ammonia and chlorides, even
at the ppm level in the fuel, degrade cell performance. These do not inhibit the platinum
catalyst per se but react with the phosphoric acid to form salts that decrease the acidity
of the electrolyte and can precipitate and block the porous electrodes. To avoid
Phosphoric Acid Fuel Cells
unacceptable performance losses, the concentration of ammonium phosphate ((NH4)H2PO4)
in the host electrolyte must be kept below 0.2 mol.%. To achieve this requirement, an
ammonia trap is usually inserted between the outlet of the fuel processor and the inlet
of the anodes to prevent ammonia from entering the stacks. The PAFC catalysts can
also degrade through the agglomeration of platinum particles. During operation, the
particles have a tendency to migrate to the surface of carbon and combine to form
larger particles, thereby decreasing the available active surface area. The rate of this
type of degradation depends mainly on the operating temperature. An unusual difficulty
is that corrosion of carbon becomes a problem at high cell voltages (above about 0.8 V).
For practical applications, low current densities with cell voltages above 0.8 V and hot
idling at open circuit are therefore best avoided with the PAFC.
7.3.3
Stack Construction
The PAFC stack consists of a repeating arrangement of a ribbed bipolar plate, the anode,
the electrolyte matrix and the cathode. In a similar manner to that described for the
PEMFC, the ribbed bipolar plate serves to separate the individual cells and to connect
them electrically in series while providing the gas supply to the anode and cathode,
respectively, as shown in Figure 1.9. As discussed previously, there is an additional
requirement in the PAFC to build in a reservoir of phosphoric acid. This feature can be
located in the electrode substrates or GDL. When made from porous graphitic carbon,
the ribbed bipolar plates can also serve as a reservoir for excess phosphoric acid. This
capability is realized in a modern PAFC stack by building a ‘multilayer’ bipolar plate, in
which a graphitic flow‐field plate is bonded on either side of a thin non‐porous carbon
layer which forms the gas barrier between adjacent cells. A generic arrangement is
illustrated in Figure 7.6 and the configuration deployed in the water‐cooled PAFC
stacks produced by International Fuel Cells (IFC), a subsidiary of United Technologies
Corporation, is shown in Figure 7.7. In the IFC design, the catalyst layers are deposited
onto porous carbon paper substrates for both cathodes and anodes. These are in turn
aec aec
Standard one piece
bipolar plate
Two-cell stack
Figure 7.6 Cell arrangement using ribbed substrates (bipolar plates): a, anode; e, electrolyte; and c,
cathode layers.
199
200
Fuel Cell Systems Explained
Oxidant gas channels
Gas-tight metal plate
Cathode catalyst layer
Electrolyte matrix
Anode catalyst layer
Fuel gas channels
Cooling water channel
Single cell
Gas flow channels
Figure 7.7 Schematic basic design of water‐cooled PAFC stacks produced by International Fuel Cells
(IFC). The figure shows a cross section through two cells of the stack. (Source: Adapted from Kurzweil,
P, 2003, Fuel Cell Technology, Vieweg, Teubner, Wiesbaden.)
bonded, using a polymer that decomposes on heating, to the flow‐field plates into which
the channels are pressed. The resulting ‘multilayer’ bipolar plate has the following
advantages over previously adopted stack configurations:
●
●
●
The surfaces between catalyst layer and GDL substrate promote uniform gas diffusion
to the electrode.
The plate is amenable to a continuous manufacturing process since the ribs on each
substrate run in only one direction; cross‐flow configuration, if required, can be easily
accommodated.
The substrate and flow‐field plates can act as a reservoir for phosphoric acid and
thereby offer a means to increase the lifetime of the stack.
A typical PAFC stack may contain 50 or more cells connected in series to obtain the
practical voltage level required.
7.3.4
Stack Cooling and Manifolding
Phosphoric acid fuel‐cell stacks can be cooled by liquid (usually water or antifreeze
solution), a dielectric (oil), or air. Cooling channels or pipes can be located between
groups of cells in the stack. As shown in Figure 7.7, cooling can also most easily be
achieved by circulating the cooling fluid between the gas‐tight components of the
bipolar plates. Note that it is not necessary for the coolant to flow between every cell—
usually, between about every fifth cell is sufficient. Air‐cooled PAFC stacks have also
been produced and offer the advantages of simplicity, reliability and low cost. The
channels in air‐cooled stacks are, however, large and this imposes a limit to the practical
size of stacks. Better heat removal is achieved with liquids that require only narrow
Phosphoric Acid Fuel Cells
channels, leading to more compact stack design. Conversely, narrow channels may be
complex to design and costly to fabricate.
Whereas small PAFC stacks may be cooled with air, stacks above about 50 kW
invariably employ either boiling or pressurized water as the coolant. With the former
method, the heat of vaporization of water is used to remove the heat from the cells.
Since the average cell temperature is around 180–200°C, the temperature of the cooling
water will be about 150–180°C. Reasonably uniform temperature in the stack can be
attained with boiling water and thereby leads to increased cell efficiency. If the
alternative of pressurized water is employed, the heat is only removed from the stack by
the heat capacity of the liquid water, so a greater flow of coolant is required. Nevertheless,
pressurized water is easier to control and, while not so efficient as boiling water, it
provides a better overall performance than that obtained with oil (dielectric) or air as
the cooling medium.
The main disadvantage of water cooling is that water treatment is necessary to
prevent the corrosion of cooling pipes and the formation of blockages in the cooling
loops. The water quality required is similar to that demanded by boilers in conventional
thermal power stations. Although not difficult to achieve with ion‐exchange resins,
such water treatment adds to the capital cost of PAFC systems.
All PAFC stacks are fitted with manifolds that are usually attached to the outside of
the stacks; these are so‐called external manifolds. (It will be noted in Section 8.4.1,
Chapter 8 that an alternative ‘internal manifold’ arrangement is preferred by some
developers of MCFC systems.) Respective inlet and outlet manifold systems enable fuel
gas and oxidant to be circulated through each cell of a particular stack. To minimize
temperature variations within the stack and thereby ensure long lifetimes, the inlet
manifold for the fuel gas is carefully designed to provide a uniform supply to each cell.
Often a stack is made of several sub‐stacks, in which the plates are mounted horizontally
on top of each other with separate fuel supplies to each sub‐stack.4 If the fuel‐cell
stack is to be operated at high pressure, the whole stack assembly has to be located
within a vessel that is filled with nitrogen gas at a pressure slightly above that of the
reactants.
7.4
Performance
The performance (voltage–current) curve for a typical PAFC is similar to that shown in
Figure 3.1 for medium to low‐temperature cells, although the current density of PAFC
stacks is usually in the range 150–400 mA cm−2. When operating at atmospheric pressure,
the output gives a cell voltage of between 600 and 800 mV. As with the PEMFC, the major
voltage losses occur at the cathode, and the overpotential is greater with air (typically
560 mV at 300 mA cm−2) than with pure oxygen (typically 480 mV at 300 mA cm−2)
because of the dilution of oxygen with nitrogen in the former. The voltage losses at the
anode are very low (ca. 4 mV per 100 mA cm−2) with pure hydrogen, which increases
4 Reference will often be found in the fuel‐cell industry to ‘sub‐stack’ and ‘short stack’. These two terms
each describe a small group of full‐size cells (i.e. the same cell area as used in completely assembled stacks).
Manufacturers routinely carry out lifetime tests on ‘short stacks’ to avoid the cost of manufacturing a full
stack of cells. The performances of a short stack and a full stack are expected to be very similar.
201
202
Fuel Cell Systems Explained
when carbon monoxide is present in the fuel gas. The ohmic loss in PAFCs is also
relatively small, namely, about 12 mV per 100 mA cm−2.
7.4.1
Operating Pressure
For any type of fuel cell, performance is a function of pressure, temperature, and
composition and utilization of the reactant gas. It is well known that an increase in the
operating pressure boosts the performance of the PAFC and, indeed, all other candidate
fuel cells. The increase in cell voltage resulting from a change in system pressure from
P1 to P2 is given by the formula (see Section 2.5.4, Chapter 2):
V
P
RT
ln 2
4F
P1
(2.44)
The change in voltage is not, however, the only benefit of a higher pressure. At the
operating temperature of the PAFC, raising the pressure also decreases the activation
overpotential at the cathode, due to the concomitant increase in the partial pressure of
both oxygen and product water. If the partial pressure of water is allowed to increase, a
lower phosphoric acid concentration will cause a slight enhancement of the ionic
conductivity that, in turn, will bring about a higher exchange-current density. This
important beneficial effect, which has been discussed in detail in Section 3.4.2,
Chapter 3, promotes further reduction of the activation overpotential, and the greater
conductivity lessens the ohmic losses. The end result is that, for the PAFC, the increase
in voltage with pressure is much higher than what is predicted by equation (2.44). From
experimental data collected over some period, the US Department of Energy Fuel Cell
Handbook5 suggests that the formula:
V
63.5 ln
P2
P1
mV
(7.7)
is a more reasonable approximation for a temperature range of 177°C < T < 218°C and a
pressure range of 0.1 MPa < P < 1.0 MPa.
7.4.2
Operating Temperature
The reversible voltage of a hydrogen fuel cell decreases as the temperature increases;
see Section 2.3, Chapter 2. Over the possible temperature range of the PAFC, the effect
is a decrease of 0.27 mV per °C under standard conditions (at which the product of
hydrogen oxidation is water vapour). On the other hand, an increase in temperature has
a beneficial effect on cell performance because activation overpotential, mass-transfer
overpotential and ohmic losses are all reduced, as discussed in Chapter 3. The kinetics
for the reduction of oxygen on platinum also improves as the cell temperature increases.
The abovementioned Fuel Cell Handbook states that for a PAFC running on air and
5 EG&G Technical Services, Inc., 2004, Fuel Cell Handbook (7th Edition), US Department of Energy, Office
of Fossil Energy, National Energy Technology Laboratory, P.O. Box 880, Morgantown, West Virginia
26507‐0880.
Phosphoric Acid Fuel Cells
750
Cell voltage/mV
700
H2
650
H2 + H2S
600
H2 + CO
SCG
550
190
170
210
230
Temperature/°C
Figure 7.8 Effect of temperature on PAFC cell voltage for different fuels: H2, H2 + 20 ppm H2S, H2 + CO
and simulated coal gas (SCG). (Source: Reproduced from Jalan, V, Poirier, J, Desai, M and Morrisean, B,
1990, Development of CO and H2S tolerant PAFC anode catalysts, Proceedings of the Second Annual
Fuel Cell Contractors Review Meeting, 2–3 May 1990, Morgantown, WV.)
pure hydrogen at a mid-range operating load (~250 mA cm−2), the voltage gain (ΔVT)
with increasing temperature is given by:
VT
1.15 T2 T1 mV
(7.8)
The data collected to derive this equation suggests that it is reasonably valid for a
temperature range of 18°C < T < 25°C. The relationship shows that each degree increase in
cell temperature should produce a performance increase of 1.15 mV. Other data indicate
that the coefficient may actually be in the range of 0.55–0.75, rather than 1.15.
Although temperature has only a minimal effect on the hydrogen oxidation reaction
at the anode, this operational parameter is important in terms of anode poisoning.
Increasing the cell temperature results in increasing anode tolerance to carbon monoxide,
as demonstrated in Figure 7.8. The benefit is a result of reduced adsorption of the gas.
A strong temperature effect for simulated coal gas (SCG) is also seen in Figure 7.8.
7.4.3
Effects of Fuel and Oxidant Composition
As mentioned in Section 7.2, fuel and oxidant utilizations are important operating
parameters for the PAFC and indeed for all types of fuel cell. In a fuel gas that is obtained,
for example, by steam reforming of natural gas, the CO2 and unreacted hydrocarbons
(e.g., CH4) are electrochemically inert and act as diluents. Because the anode reaction is
nearly reversible, the fuel composition and hydrogen utilization generally do not
strongly influence cell performance. Cell voltage will, however, be influenced by a
change in the partial pressure of hydrogen that can result from a change in either the
composition or the utilization of the fuel. This effect can be described by a relationship
similar to equation (7.7), namely:
V
55 ln
P2
P1
mV
(7.9)
203
204
Fuel Cell Systems Explained
On the cathode side, the use of air with ~21 vol.% oxygen instead of pure oxygen
results in a decrease in the current density by a factor of about three at constant
electrode potential. The overpotential at the cathode increases with an increasing
consumption of oxygen.
7.4.4
Effects of Carbon Monoxide and Sulfur
It has been stated already that platinum in the anode catalyst of the PAFC may be
poisoned by carbon monoxide in the fuel gas. At low concentrations of CO, absorption
on the anode electrocatalyst is reversible, and CO will be desorbed if the temperature
is raised. The methods undertaken to limit the CO concentration are discussed in
Section 10.4.11, Chapter 10.
Sulfur in the fuel stream, usually present as hydrogen sulfide (H2S), will similarly
poison the anode of a PAFC. State‐of‐the‐art PAFC stacks are able to tolerate up to
50 ppm of sulfur in the fuel. Sulfur poisoning does not affect the cathode, and moderately
poisoned anodes can be reactivated by increasing the temperature.
7.5 Technological Developments
Until recently, the PAFC was the only fuel‐cell technology that could be said to be
available commercially. Systems are ready to meet market specifications and are
supplied with guarantees. Many of the systems built by IFC have now run for several
years, and so there is a wealth of operating experience from which developers and endusers can draw. One important aspect that has arisen from field trials of the early PAFC
plants is the reliability of the stack and the quality of power produced by the systems.
These dual attributes have led to systems being preferred for so‐called ‘premium power’
applications, such as in banks, hospitals and computing facilities. Worldwide, PAFC
plants with a total installed capacity in excess of 65 MW have been tested, are being
tested, or are being fabricated. Most of the systems are in the capacity range of
50–200 kW, but large versions of 1 and 5 MW have also been constructed. The largest
plant operated to date has been that built by IFC and Toshiba for Tokyo Electric Power.
This facility can generate 11 MW of grid‐quality AC power. Efforts in the United States
and Japan are now concentrating on the improvement of PAFCs for stationary dispersed
power and on‐site cogeneration (CHP). The major industrial developers are Doosan
Fuel Cell America Inc. in the United States and Fuji Electric, Toshiba and Mitsubishi
Electric Corporation in Japan.
Although the PAFC has now reached a level of maturity where customer confidence
can be guaranteed, the technology is still too costly to be economic compared with
alternative power‐generation systems, except in the niche premium power applications
referred to in the previous text. There is a need to increase the power density of the cells
and reduce capital costs; both issues are inextricably linked. System optimization is also
a key issue. Much of the recent development in the technology is proprietary, but the
following overview gives an indication of progress made during the past few years.
During the early 1990s, the goal of the research and development of PAFCs in the
United States was to design and demonstrate a large stack with a power density of
0.188 W cm−2, a practical life of 40 000 h and a stack cost of less than US$400 per kW.
Phosphoric Acid Fuel Cells
A conceptual design of an improved technology stack operating at 820 kPa and 200°C
was produced. The stack would be composed of 355 × 1 m2 cells to produce over
1 MW of DC power in the same physical envelope as the unit 670‐kW stack for
the 11‐MW PAFC plant built for Tokyo Electric Power. The improvements made to
the design were tested in single cells, as well as in sub‐scale and full‐size short stacks.
The results of these tests were outstanding. The power density goal was exceeded;
namely, 0.323 W cm−2 was achieved in single cells when operating at 645 mA cm−2 and
up to 0.66 V per cell. A cell performance of 0.307 W cm−2 was obtained from stacks,
with an average of 0.71 V per cell at 431 mA cm−2. By comparison, in 1991 the 11‐MW
Tokyo Electric Power’s system gave an average cell performance of approximately
0.75 V per cell at 190 mA cm−2. The rate of performance degradation of the stacks was
less than 4 mV per 1000 h during a test of 4500 h. The results from this programme
represent the highest performance of full‐size phosphoric acid cells and short stacks
published to date.
Mitsubishi Electric Corporation has also demonstrated an enhanced performance of
0.65 V at 300 mA cm−2 in single cells. Component improvements by Mitsubishi
have resulted in the lowest rate of PAFC degradation to be publicly acknowledged,
namely, 2 mV per 1000 h for 10 000 h at 200–250 mA cm−2 in a short stack with cells of
3600 cm2 area.
Catalyst development continues to be an important aspect of the future of the PAFC.
Over the past 10 years, several non‐precious metal catalysts for cathodes have been
investigated and are similar to those for PEMFC catalysts, as described in Section 4.3.2,
Chapter 4. These include transition metal (e.g., iron or cobalt) organic macrocycles
of tetramethoxyphenylporphyrin (TMPP), phthalocyanines (PC), tetraazaannulene
(TAA) and tetraphenylporphyrin (TPP). Another approach has been to alloy platinum
with transition metals such as nickel, titanium, chromium, vanadium, zirconium and
tantalum. Notable work by Johnson Matthey during the early 2000s showed that
platinum–nickel alloy catalysts yielded a 49 wt.% increase in specific activity over pure
platinum. This advance is translated into a 39 mV improvement in the performance of
the air electrode at 200 mA cm−2.
Other recent significant advances in PAFC technology are improved construction of
the gas‐diffusion electrode and materials that offer greater protection against carbon
corrosion. Of course, there is scope for many changes in the system design, with better
engineered balance‐of‐plant components such as the reformer, shift reactors, heatexchangers and burners. Power electronics for DC to DC or DC to AC conversion have
improved significantly since early units built in the 1990s, in terms of both size and
performance (see Section 12.2.1, Chapter 12). Nonetheless, in the 400‐kW systems now
supplied by Doosan, the footprint and weight of the fuel processing components are
significant in comparison with the stack modules.
Some of the earliest PAFC system demonstrations were conducted in the 1970s under
the ‘Target Program’ funded by the American Gas Association. Many organizations
have been involved in the development of PAFCs, but in recent years the thrust of
commercialization has been borne mainly by two companies, namely, UTC Fuel Cells
(formerly trading under the name ONSI or International Fuel Cells) based in
Connecticut, USA, and Fuji Electric in Japan. The latter has been developing PAFCs
since the 1980s and in the 1990s started to supply 50‐kW and 100‐kW systems
worldwide. Over 100 such systems have been commissioned—a testament to their
205
206
Fuel Cell Systems Explained
reliability and durability. Current research at Fuji is directed towards raising performance
through better reforming catalysts and reducing costs, especially with respect to the
balance‐of‐plant equipment.
UTC Fuel Cells has supplied globally several hundred model PC25 200‐kW systems.
The technology has found a niche application in high‐value locations such as a post
office facility in Alaska, a science centre in Japan, the New York City Police Department
and the First National Bank of Omaha. The last‐mentioned application is especially
interesting as several fuel cells are linked together with other generation equipment to
create an ultra‐reliable power system to sustain a critical load. Sure Power Corporation
who put together the equipment has guaranteed 99.9999% reliability of the whole power
system, thereby capitalizing on the reliability and robustness of PAFC technology. There
are no single points of failure,6 and the Sure Power product is extremely fault tolerant.
In 1987, Bharat Heavy Electricals Ltd. of India started to fund the research and
development of PAFC systems. In 2001, the company had built a 50‐kW prototype, but
shortly afterwards the company cut back its effort on PAFCs in favour of PEMFCs.
Similarly, Caltex Oil Corporation in South Korea engaged in the construction of a
50‐kW system during the 1990s. This undertaking, too, does not appear to have been
taken further. The Japanese companies Sanyo, Toshiba and Mitsubishi Electric all
produced PAFC stacks during the 1980s and early 1990s, but there is little evidence
that these companies have continued their efforts on PAFCs. As technical progress in
PEMFC technology has moved forward, many organizations once involved in PAFCs
have shelved their activities or used them to enhance their own PEMFC development.
Further Reading
Behling, N, 2012, History of phosphoric acid fuel cells, in Fuel Cells: Current Technology
Challenges and Future Research Needs, pp. 53–135, Elsevier, Amsterdam, the
Netherlands.
Sammes, N, Bove, R and Stahl, K, 2004, Phosphoric Acid Fuel Cells: Fundamentals and
Applications, Current Opinion in Solid State and Materials Science, vol. 8(5),
pp 372–378.
6 A single point of failure (SPOF) is a part of a system that, if it fails, will stop the entire system from
working. SPOFs are undesirable in any system with a goal of high availability or reliability, be it a business
practice, software application or other industrial systems.
207
8
Molten Carbonate Fuel Cells
8.1
Principles of Operation
The electrolyte of the molten carbonate fuel cell (MCFC) is a molten mixture of alkali
metal carbonates — usually a binary mixture of lithium and potassium, or lithium and
sodium carbonates — which is retained in a ceramic matrix of lithium aluminate (LiAlO2).
At the high operating temperatures (typically 600–700°C), the alkali carbonates form a
highly conductive molten salt, with carbonate CO32− ions providing ionic conduction.
The anode and cathode reactions are shown schematically in Figure 8.1. Note that,
unlike other common types of fuel cell, carbon dioxide (CO2) has to be supplied to the
cathode as well as oxygen, and this becomes converted to carbonate ions, which migrate
to the anode where reconversion to CO2 occurs. For every mole of hydrogen that is
oxidized in the cell, there is therefore a net transfer of one mole of CO2 along with two
Faradays of charge or two moles of electrons between the two electrodes. Note that
the requirement for CO2 to be supplied to the MCFC contrasts with the alkaline fuel
cell (AFC) from which CO2 must be excluded. The overall reaction of the MCFC is
therefore:
H2
1
O2
2
CO2 cathode
H2 O CO2 anode
(8.1)
The Nernst reversible voltage for an MCFC, taking into account the transfer of CO2,
is given by the equation
1
Vr
Vr
RT
ln
2F
PH2 . PO22
PH2 O
PCO2c
RT
ln
2F
PCO2a
(8.2)
where sub‐subscripts a and c refer to the anode and cathode gas compartments,
respectively. Usually, there is a difference in the partial pressures of CO2 between the
two electrodes, but when these pressures are identical, the cell potential depends only
on the partial pressures of H2, O2 and H2O.
In an MCFC system, it is normal practice to recycle externally the CO2 generated at
the cell anodes to the cathodes where it is consumed. Whereas the recycling might at
first seem to be an added complication and therefore place the MCFC at a disadvantage
Fuel Cell Systems Explained, Third Edition. Andrew L. Dicks and David A. J. Rand.
© 2018 John Wiley & Sons Ltd. Published 2018 by John Wiley & Sons Ltd.
208
Fuel Cell Systems Explained
Hydrogen fuel
Anode 2H2 + 2CO32– → 2H2O + 2CO2 + 4e–
CO32– ions through electrolyte
Cathode O2 + 2CO2 + 4e–
2CO32–
Load
(e.g., electric
motor)
Electrons flow round
the external circuit
Oxygen and carbon dioxide
Figure 8.1 Anode and cathode reactions for a MCFC using hydrogen fuel. Note that the product
water is at the anode and that both carbon dioxide and oxygen have to be supplied to the cathode.
CH4
CO2 + CH4 + steam
Anode
Fuel in
Electrolyte
Cool air
Cathode
Hot air + CO2
+ steam
Burner
Hot oxygen depleted air
+ CO2 + steam
Figure 8.2 Adding carbon dioxide to the cathode gas stream does not add complexity.
compared with other fuel‐cell types, it can be achieved by feeding the anode exhaust gas
to a combustor (burner), which converts any unused hydrogen or fuel gas (e.g., methane,
CH4) into water and CO2. The exhaust gas from the combustor is then mixed with fresh
air and fed to the cathode inlet, as shown in Figure 8.2. The process is no more complex
than that for other types of high‐temperature fuel cell, and, moreover, it also serves to
preheat the reactant air, to burn the unused fuel and to bring the waste heat into one
stream for use in a bottoming cycle or for other purposes.
Another possible method of supplying CO2 to the cathode inlet is to use a device,
such as a membrane, that will separate CO2 from the anode exit gas and allow it to be
transferred to the cathode inlet gas. The advantage of using such a ‘transfer device’ is
that any unused fuel gas can be recycled to the anode inlet or used for other operations
such as providing heat for fuel processing. A further alternative would be to supply CO2
from an external source, especially where a ready supply of the gas is available.
At the operating temperature of MCFCs, nickel (anode) and nickel oxide (cathode),
respectively, are adequate catalysts to promote the two electrochemical reactions.
Unlike the phosphoric acid fuel cell (PAFC) or the proton‐exchange membrane fuel cell
Molten Carbonate Fuel Cells
(PEMFC), noble metals are not required. Other important differences between the
MCFC and the PAFC or PEMFC are its ability to perform direct electrochemical
conversion of carbon monoxide (CO) and to reform hydrocarbon fuels internally. If CO
were to be fed as fuel to the MCFC, the reactions at each electrode given in Figure 8.3
would occur.
The voltage of the fuel cell when operating with CO is calculated in exactly the same
way as for the hydrogen fuel cell, as described in the Section 2.1, Chapter 2. Two
electrons are released for each molecule of CO, as illustrated in Figure 8.3, just as two
electrons are released for each molecule of H2. Consequently, the formula for the open‐
circuit voltage is identical, i.e.,
Vr
gf
(8.3)
2F
The method of calculating ∆ g f is given in Appendix 1. Notably, the values for hydrogen
and carbon monoxide are remarkably similar at 650°C, as shown in Table 8.1.
In practical applications, it is most unlikely that pure CO would be used as a fuel. It is
more practical for the fuel gas to contain both H2O and CO, and in such cases the
electrochemical oxidation of the CO would probably proceed via the water-gas shift
reaction, i.e., equation (7.3), Chapter 7, which is a fast reaction that occurs on the nickel
electrocatalyst of the anode. The shift reaction converts CO and steam to hydrogen that
then oxidizes rapidly. The two reactions — direct oxidation of CO or shift reaction and
then the oxidation of H2 — are entirely equivalent.
Carbon monoxide fuel
Anode
2CO + 2CO32– → 4CO2 + 4e–
Load
(e.g., electric
motor
CO32– ions through electrolyte
Cathode O2 + 2CO2 + 4e–
2CO32–
Electrons flow round
the external circuit
Oxygen and carbon dioxide
Figure 8.3 Anode and cathode reactions for an MCFC using carbon monoxide fuel.
Table 8.1 Values of ∆ g fo and Vro for hydrogen and carbon
monoxide fuel cells at 650°C.
Fuel
∆ g fo (kJ mol−1)
Vro (V)
H2
–197
1.02
CO
–201
1.04
209
210
Fuel Cell Systems Explained
Unlike the PEMFC, AFC and PAFC, the MCFC operates at a temperature that is
sufficiently high to enable internal reforming of hydrocarbon fuels such as methane.
This is a particular strong feature of both the MCFC and, as shall be discussed later in
Chapter 9, the solid oxide fuel cell (SOFC). In internal reforming, steam is added to
the fuel gas before it enters the stack. Inside the stack, the fuel and steam react in the
presence of a suitable catalyst according to steam reforming reactions such as those
given by equations (7.1) and (7.2), Chapter 7. Heat for the endothermic reforming
reactions is supplied by the electrochemical reactions of the cell. Internal reforming is
discussed in more detail in Chapter 10.
Compared with low‐temperature fuel cells, the high operating temperature of the MCFC
provides the opportunity for achieving superior overall system efficiencies and greater
flexibility in the use of available fuels. Unfortunately, however, the higher temperature
places severe demands on the corrosion resistance and life of cell components in the
aggressive environment of the molten carbonate electrolyte.
The PAFC and MCFC are similar in that they both use a liquid electrolyte that is
immobilized within a porous solid matrix. As discussed in Section 7.3.2, Chapter 7,
PTFE employed in a PAFC serves as both a binder and a wet‐proofing agent to maintain
the integrity of the electrode structure and to establish a stable electrolyte|gas interface
in the porous electrodes. The phosphoric acid is retained in a matrix of SiC doped with
PTFE that is sandwiched between the anode and cathode. Since there are no materials
comparable with PTFE that are able to withstand MCFC temperatures, a different
approach is necessary to establish a stable electrolyte|gas interface in the porous
electrodes of the MCFC. Stable interfacial boundaries have been achieved through a
balance in capillary pressures.
The porous electrolyte matrix generally has a narrow pore‐size distribution of
relatively small pores. By contrast, the electrodes are characterized by a much
wider pore‐size distribution of pores with much larger diameters. These different
characteristics between electrolyte and electrodes allow the electrolyte matrix to
remain completely filled with molten carbonate, whereas the porous electrodes are
only partially filled. The pores in the electrodes will be partially filled in inverse
proportion to the pore size, namely, the larger the pore, the less they are filled.
Electrolyte management, i.e., the control over the optimum distribution of electrolyte
in the different cell components, is critical for achieving high performance and good
service life from MCFCs. This operational feature is extremely specific to MCFCs,
and it should be emphasized that the use of a liquid electrolyte constrained within
the matrix by capillary forces can give rise to various undesirable processes. Examples
are the consumption of electrolyte by corrosion reactions, potential‐driven migration
or creepage of liquid electrolyte and vapourization of carbonates or hydroxide
species, all of which can contribute to the redistribution or loss of molten carbonate
from the cells.
8.2
Cell Components
In the early days of MCFC development, precious metals were generally used as the
electrode materials. During the 1960s and 1970s, however, nickel‐based alloys became
the preferred choice for the anode and nickel oxide for the cathode. Since that time,
Molten Carbonate Fuel Cells
Table 8.2 Evolution of cell component technology for molten carbonate fuel cells.
Component
~1965
~1975
Current status
Anode
Pt, Pd or Ni
Ni ‐ 10 wt.% Cr
Ni–Cr or Ni–Al
3–6 µm pore size
45–70% initial porosity
0.20–1.5 mm thickness
0.1–1 m2 g−1
Cathode
Ag2O or lithiated NiO
Lithiated NiO
Lithiated NiO
7–15 µm pore size
70–80% initial porosity
60–65% after lithiation and
oxidation
0.5–1 mm thickness
0.5 m2 g−1
Electrolyte
support
MgO
Mixture of
α‐, β‐, γ‐LiAlO2
10–20 m2 g−1
Electrolyte
a
α‐LiAlO2 β‐LiAlO2
0.1–12 m2 g−1
0.5–1 mm thickness
52 Li–48 Na
62 Li–38 K
62 Li–38 K
43.5 Li–31.5 Na–25 K
~60–65 wt.%
50 Li–50 Na
Hot press ‘tile’
~50 wt.%
1.8 mm thickness
Tape cast
‘Paste’
0.5–1 mm thickness
Source: Hirschenhofer, JJ, Stauffer, DB and Engelman, RR, 1998, Fuel Cell Handbook, 4th edition, Report No.
DOE–FETC‐99/1076, Parsons Corporation, for U.S. Department of Energy.
a) Figures in this entry are in mol.% unless stated otherwise.
both the electrode materials and the electrolyte structure (molten carbonate in a
ceramic matrix of LiAlO2) have remained essentially unchanged. On the other hand,
from the 1980s onwards, there has been an evolution in the methods employed for
fabricating the electrolyte structures. Some of the principal materials used in MCFCs
since the 1960s are summarized in Table 8.2.
8.2.1
Electrolyte
The MCFC electrolyte is generally a mixture of alkali metal carbonates. Lithium offers
the best ionic conductivity due to its low atomic weight, and in the beginning eutectic
mixtures of Li–Na carbonate (Li2CO3–Na2CO3) were used; then attention shifted to
Li–K (Li2CO2–K2CO3), especially the eutectic mixture of 62 : 38 mol.% Li2CO2–K2CO3,
which proved effective for atmospheric pressure operation due to its increased reactivity.
Ternary mixtures of Li–K–Na have also been investigated, but the present trend is to go
back to mixtures of Li–Na, due to their lower vapour pressure that reduces electrolyte
loss and their increased basicity compared with Li–K mixtures.
211
212
Fuel Cell Systems Explained
The molten electrolyte is constrained in a ceramic fibrous matrix of high surface area
submicron lithium aluminate (LiAlO2). Historically, the aluminate has been prepared
from the low‐temperature γ form of alumina, and developers are comfortable in using
this for the preferred tape‐casting manufacturing process. The matrix may be enhanced
by the addition of particles of α‐alumina or zirconia to improve its mechanical strength.
The material appears to be reasonably stable when supporting electrolytes based on
lithium and potassium carbonate. If sodium carbonate is employed, e.g., as a Li–Na
mixture, however, the particle size of γ‐LiAlO2 has been shown to increase over a period
of time along with a phase change to α‐LiAlO2. The reverse phase change has also been
noted in systems where α‐LiAlO2 has been employed with Li–K carbonate mixtures.
Such phase changes can lead to a loss of structural integrity within the matrix, giving
rise to cracks and cell degradation. No universal solution to this problem has yet been
identified, although some improvements in the mechanical strength of the matrix have
been reported by adding particles of alumina or even micro‐sized particles of aluminium
to the slurry before it is tape‐cast.1 Until the 1990s, the matrix was often fabricated
into a tile by hot‐pressing the powdered material, and it is still often referred to as the
electrolyte ‘tile’. Nowadays, the matrix is invariably made using tape‐casting methods
that are commonly employed in the ceramics and electronics industries. The basic
process involves dispersing the ceramic materials in a ‘solvent’ that contains dissolved
binders (typically, organic compounds), plasticizers and additives to achieve the desired
viscosity and rheology of the resulting mixture or ‘slip’. The slip material is then cast in
the form of a thin film over a moving smooth surface, and the required thickness is
obtained by shearing with an adjustable blade (a so‐called ‘doctor blade’ assembly).
After drying, the film is heated further in air and any organic binder is burnt off at
250–300°C. The semi‐stiff ‘green’ product is then assembled into the stack structure.
Tape casting of the electrolyte provides an effective means of fabricating components of
large area. The methods described here can also be applied to the cathode and anode
materials for the ready manufacture of stacks with an electrode area up to about 1 m2.
Compared with most other fuel cells, the ohmic resistance of the MCFC electrolyte,
and especially the ceramic matrix, has a major influence on the operating voltage.
Under typical operating conditions, the electrolyte has been found to account for 70%
of the ohmic losses in a MCFC.2 Furthermore, the losses are dependent on the thickness
of electrolyte according to the formula:
Vr
0.533t
(8.4)
where t is the thickness in cm. The relationship shows that a fuel cell with an electrolyte
structure of 0.025 cm thickness would operate at a cell voltage that is 82 mV higher than
an identical cell with a structure of 0.18 cm thickness. Using tape‐casting methods,
electrolyte matrices can now be made quite thin (0.25–0.5 mm), which is a development
that offers a significant advantage in reducing the ohmic resistance. There is a trade‐off,
1 Kim, J, Patil, K, Han, J, Yoon, SP, Nam, S, Lim, T, Hong, S, Kim, H and Lim, H, 2009, Using aluminum and
Li2CO3 particles to reinforce the α‐LiAlO2 matrix for molten carbonate fuel cells, International Journal of
Hydrogen Energy, vol. 34, pp. 9227–9232.
2 Yuh, C and Farooque, JR, 1992, Understanding of carbonate fuel cell resistances in MCFCs. Proceedings
of the Fourth Annual Fuel Cell Contractors Review meeting, U.S. DOE–METC, pp. 53–57.
Molten Carbonate Fuel Cells
however, between low resistance and the long‐term stability that is otherwise obtained
with thicker materials.
With respect to the electrolyte, there is an important difference between the MCFC
and all other types of fuel cell. That is, the final preparation of the cell is carried out once
the stack components are assembled. Layers of electrodes, electrolyte and matrix,
together with various non‐porous components (current-collectors and bipolar plates),
are assembled together, and the whole package is heated slowly up to the operating
temperature of the fuel cell. As the carbonate reaches its melt temperature (over 450°C),
it becomes absorbed into the ceramic matrix. This process results in a significant
shrinkage of the stack that must be accommodated by careful mechanical design of the
total assembly. During the heating, a reducing gas has to be supplied to the anode side
of the cell to ensure that the nickel anode remains in the chemically reduced state.
8.2.2
Anode
State‐of‐the‐art MCFC anodes are made of porous sintered nickel alloyed with a small
amount of chromium and/or aluminium (see Table 8.2). The flat planar anode usually
has a thickness of 0.4–0.8 mm, a porosity of between 55 and 75% and a mean pore
diameter of 4–6 µm. Fabrication involves either hot‐pressing finely divided powder or
tape‐casting a slurry of the powdered material mixed with additives to achieve the
desired fluid properties of the powder–binder mixture. Tape casting is a low‐cost wet
process that produces a thinner anode with more control over the thickness and pore‐
size distribution than the alternative procedure of hot pressing a powder. The addition
of chromium or aluminium (usually 10–20 wt.%) improves the mechanical stability of
the porous nickel by reducing the sintering of nickel particles during cell operation.
Unless controlled, sintering can become a major problem in the MCFC anode because
it leads to growth in pore size, a reduction in surface area and a loss of carbonate from the
electrolyte. The change in pore structure can also result in mechanical deformation of
the anode under the compressive load in the stack that, in turn, decreases electrochemical
performance and may cause cracking of the electrolyte.
Although Ni–Cr or Ni–Al alloy anodes have achieved acceptable stability for
commercial applications, the cost is relatively high, and consequently developers have
been investigating alternative materials. Partial substitution of the nickel with copper, for
example, can go some way to reducing the cost of the alloys, but complete substitution
is not feasible as copper exhibits more creep than nickel. In an attempt to improve the
tolerance to sulfur in the fuel stream, various ceramic anodes are also being investigated.
These include LiFeO2 with and without the doping of manganese or niobium.
The anode of the MCFC has to provide more than just electrocatalytic activity.
Because the anode reaction is relatively fast at MCFC temperatures, a high surface area
is not required, compared with the cathode. Partial flooding of the anode with molten
carbonate is therefore acceptable, and this feature is used to good effect, not only to act
as a reservoir for carbonate (i.e., much in the same way that the porous carbon substrate
functions in the PAFC) but also to replenish carbonate that may be lost from a stack
during prolonged use.
In some of the earlier MCFC stacks, a so‐called bubble barrier was located between
the anode and the electrolyte. The bubble barrier consisted of a thin layer of Ni or
LiAlO2 that had only small pores. The component served to prevent a flow of electrolyte
213
214
Fuel Cell Systems Explained
to the anode, as well as to lower the risk of gas crossover. As discussed earlier, this latter
problem is common to all liquid fuel cells in which an excess of pressure on one side of
the cell may cause the fuel gas to cross the electrolyte. Nowadays, the use of a tape‐cast
structure enables control of the pore distribution in anode materials during manufacture so that small pores are found closest to the electrolyte and larger pores are nearer
to the gas channels. Long‐term electrolyte loss is, however, still a significant problem
with the MCFC, and a totally satisfactory approach to electrolyte management is yet to
be achieved.
8.2.3
Cathode
Pure nickel oxide (NiO), which is an n‐type semiconductor, is the preferred choice for
the cathode material. In the MCFC environment, the oxide becomes doped with Li+
ions from the electrolyte with the concomitant creation of extra electron–hole pairs for
conduction by replacing Ni2+ with Ni3+ that, in turn, enhances the electrical conductivity.
One of the major difficulties in employing NiO for the MCFC cathode, however, is that
it has a small, but significant, solubility in molten carbonates. Consequently, some
nickel ions are formed in the electrolyte and tend to diffuse towards the anode. As the
ions approach the chemically reducing conditions at the anode (note: hydrogen is
present from the fuel gas), metallic nickel can precipitate out in the electrolyte and
cause internal short circuits with subsequent loss in the power output of the cell.
Furthermore, the deposited nickel can act as a sink for the ions and thereby promote the
further dissolution of the metal from the cathode. The leaching of nickel is intensified
at high partial pressures of CO2 through the following reaction:
NiO CO2
Ni 2
CO32
(8.5)
The problem is mitigated if the more basic, rather than acidic, carbonates are used in
the electrolyte. The basicity of the common alkali metal carbonates decreases in the
order: (basic) Li2CO3 > Na2CO3 > K2CO3 (acidic). The lowest nickel oxide dissolution
rates have been found for the eutectic mixtures of 62 wt.% Li2CO3 + 38 wt.% K2CO3 and
52 wt.% Li2CO3 + 48 wt.% Na2CO3. The addition of some alkaline earth oxides (CaO,
SrO, BaO) to the carbonates has been shown to reduce the solubility of NiO by up to
50%. It also has been reported that lanthanum oxide3 can reduce the solubility still
further, and this is thought to be due to the formation of oxy‐carbonates such as
La2O2CO3 that increase the basicity of the electrolyte melt. Such rare earth oxide
addition may also improve the rate of the oxygen reduction reaction in the MCFC; it has
been observed that with addition of 0.5 wt.% CeO2 and 0.5 wt.% La2O3, the charge‐
transfer resistance in Li–K carbonate melt decreases by an order of magnitude.4
With state‐of‐the‐art NiO cathodes, nickel dissolution can be minimized by (i) using
a basic carbonate, (ii) operating at atmospheric pressure and keeping the CO2 partial
pressure in the cathode compartment low and (iii) employing a relatively thick electrolyte
matrix to increase the Ni2+ diffusion path. By these means, cell lifetimes in excess of
3 Ota, KI, Matsuda, Y, Matsuzawa, K, Mitsushura, S and Karnia, N, 2006, Effect of rare earth oxides for
improvement of MCFC, Journal of Power Sources, vol. 160(2), pp. 811–815.
4 Scaccia, S, Frangini, S, Dellepiane, S, 2008, Enhanced oxygen solubility by Re2O3, Journal of Molecular
Liquids, vol. 138, pp. 107–112.
Molten Carbonate Fuel Cells
40 000 h have been demonstrated under atmospheric pressure operation. For operation
at higher pressure, alternative cathode materials have been investigated. LiCoO2 and
LiFeO2 have attracted the most attention, and the former has the lower dissolution rate,
which is an order of magnitude lower than NiO at atmospheric pressure. Dissolution of
LiCoO2 also shows a lower dependency on CO2 partial pressure than NiO.
Initial work on LiCoO2 in the early 1990s focused on using it simply as an alternative
to NiO. Unfortunately, the relatively higher cost of cobalt and the lower mechanical
strength of LiCoO2 compared with NiO discouraged developers to adopt it as a single‐
phase replacement. More success was obtained when LiCoO2 was combined with NiO
or when both LiCoO2 and LiFeO2 were combined with NiO. Many cathodes made with
NiO particles coated with oxides (i.e., core–shell structure) or with oxides finely
dispersed on NiO particles have been evaluated in recent years. It has been shown, for
example, that finely dispersed Ce and Co on NiO particles prepared via a polymeric
precursor route significantly reduces NiO dissolution over the short term (up to a few
hundred hours).5 It remains to be seen, in commercially operating stacks, whether
nickel oxide dissolution from Ce–Co–NiO cathodes becomes a dominating factor in
cell degradation in the longer term.
8.2.4
Non‐Porous Components
The bipolar plates for the MCFC are usually fabricated from thin sheets of stainless
steel. The anode side of the plate is coated with nickel. The coating is stable in the
reducing environment of the anode, provides a conducting path for current collection
and is not wetted by electrolyte that may migrate from the anode. Gas‐tight sealing of
the cell is achieved by allowing the electrolyte from the matrix to contact the bipolar
plate at the edge of each cell outside the electrochemically active area, as illustrated
schematically in Figure 8.4.
To avoid corrosion of the stainless steel in this ‘wet‐seal’ area, the bipolar plate is
coated with a thin layer of aluminium, which forms a protective layer of γ‐LiAlO2
through reaction with Li2CO3 in the electrolyte. There are many designs of bipolar
plate, as determined by whether the gases are manifolded externally or internally. Some
of the bipolar plates developed for internal reforming have the reforming catalyst
incorporated within the anode gas flow-field (see Figure 10.4, Chapter 10).
8.3
Stack Configuration and Sealing
The stack configuration for the MCFC is very different from those described in previous
chapters for the PEM, AFC and PAFC although there are, of course, some similarities.
The most important difference is in the method of sealing. As described in the previous
section, the MCFC stack is composed of various porous components (matrix and
electrodes) and non‐porous components (current-collector and bipolar plate). In assembling and sealing these components, it is essential to ensure an equitable distribution
of gas flows between individual cells, uniform distribution within each cell and good
5 Kim, MH, Hong MZ, Kim, YS, Park, E, Lee, H and Ha, W, 2006, Cobalt and cerium coated Ni powder as a
new candidate cathode material for MCFC, Electrochimica Acta, vol. 51, pp. 6145–6151.
215
216
Fuel Cell Systems Explained
Bipolar plate
Cathode
currentcollector
Cathode
Matrix
Wet seal
Anode
Anode currentcollector
Figure 8.4 Schematic cross-section of an MCFC: cell components exposed to high‐temperature hot
corrosion environment showing the location of the wet seals on the anode and cathode sides of the
electrolyte support matrix.
thermal management to reduce temperature gradients throughout the stack. Whereas
several proprietary techniques have been developed for constructing stacks, some
generic aspects are described in the following text, with examples from practical
systems.
8.3.1
Manifolding
Reactant gases have to be supplied in parallel to all cells in the same stack via common
manifolds. The basic arrangement for external manifolding is shown in Figure 1.11,
Chapter 1. The electrodes are about the same area as the bipolar plates, and the reactant
gases are fed into, and removed from, the appropriate faces of the fuel‐cell stack. One
advantage of external manifolding is its simplicity in enabling a low‐pressure drop in the
manifold and a good flow distribution between cells. A disadvantage is that the two gas
flows are at right angles to each other, i.e., there is a cross‐flow, and this can cause
uneven temperature distribution over the faces of the electrodes. Other problems have
been gas leakage and migration (‘ion pumping’) of the electrolyte. Each external manifold
must have an insulating gasket to form a seal with the edges of the stack. This is usually
made from zirconia felt, which provides a small amount of elasticity to ensure a good
seal. Note that most stack developers arrange the cells to lie horizontally so that the fuel
and the oxidant are supplied to the vertical sides of the stack in a cross‐flow arrangement.
The ‘Hot Module’ is an alternative arrangement pioneered by MTU Friedrichshafen
and has vertically mounted cells with the anode inlet manifold located underneath the
stack. In this way, sealing with the gasket located at the anode inlet is enhanced by the
weight of the whole stack.
Internal manifolding refers to a means of gas distribution via ducts that penetrate
through the cells within the stack. The arrangement is illustrated in Figure 1.12,
Chapter 1. An important advantage of internal manifolding is the greater flexibility in
Molten Carbonate Fuel Cells
the direction of flow of the gases. For even temperature distribution, co‐flow or counter‐
flow can be used, as discussed in Section 7.2.2, Chapter 7. The ducts that constitute the
internal manifold in an MCFC stack are formed by holes in each separator plate that
line up with each other once the stack components are assembled. The bipolar separator
plates in an MCFC stack can be quite complex mechanically as illustrated by the
IMHEX design shown schematically in Figure 8.5b and c. The parallel arrangement of
channels in this design allows for either co‐flow or counter‐flow gas configurations.
The IMHEX bipolar plate consists of several thin sheets of metal. Two of the sheets are
corrugated to form flow-fields. These sit above and below a plane solid sheet of stainless
steel that separates the fuel and oxidant gases from adjacent cells. Above and below the
corrugated sheets are perforated current‐collector plates, and above and below these
are plates that serve to act as the holders for the anode and cathode components of
adjacent cells. The electrolyte matrix extends out to the extremities of the separator
plates, and all of the plates have holes for the manifold gas ducts that align when the
Straight rib
(a)
Active area
Thickness: 0.020″
Pitch: ~0.200″
Height: ~0.100″
(b)
2
3
Current-collector
Gas inlet/outlet
Anode
Electrolyte plate
1
Cathode
Current-collector
3
2
Separator
Gas channel
(c)
Electrolyte plate
Cathode
Separator
Corrugated plate
Anode
Manifold
Perforated plate
Fuel
Figure 8.5 Examples of practical separator-plate designs with internal manifolding: (a) IMHEX design
of ECN, (b) multiple‐cell stack of Hitachi and (c) cross-section of wet‐seal area in an internally
manifolded MCFC stack.
217
218
Fuel Cell Systems Explained
stack is assembled. Thus the electrolyte itself, when molten, acts as the means of sealing
both around the internal gas ducts and around the perimeter of each cell. Using the
electrolyte in this way creates a ‘wet seal’ that prevents leakage or cross‐contamination
of gases so long as the pressure of gas inside the stack is close to that of the external
atmosphere. The advantages of using internal manifolding are therefore offset to some
extent by the more complex design of bipolar plate compared with that required for an
externally-manifolded stack.
8.3.2
Internal and External Reforming
Internal reforming has been championed in the MCFC from the early 1960s. If a
mixture of methane and steam (2 : 1 by volume) is reformed at the normal 650°C
operating temperature of the MCFC and the product gas reaches thermodynamic
equilibrium, then typically the methane conversion is about 85%. This conversion can
be obtained via indirect internal-reforming (IIR), namely, by simply inserting reforming
plates between groups of cells. Each reforming plate supports a conventional metal
catalyst. By contrast, direct internal reforming (DIR) is achieved by inserting supported
metal catalyst particles within the anode compartments of cells, i.e., within the flow‐
field channels of the bipolar-plates on the anode side. Direct internal reforming achieves
100% conversion of methane and much better heat utilization. Perhaps the most
successful application has been the combination of both IIR and DIR technologies, as
illustrated schematically in Figure 8.6. This approach is used in the Direct FuelCell™
products that are available from Fuel Cell Energy Inc. (FCE) in the United States.
Oxidant
Fuel
DIR
catalyst
Natural
gas
nit
ru
e
orm
f
Re
e
ag
Partially
reformed
fuel
k
ac
ll p
IR
ce
D
Oxidant
IIR
catalyst
Figure 8.6 Combination of IIR and DIR achieves very high electrical conversion efficiency and high
conversion of fuel.
Molten Carbonate Fuel Cells
Obviously, internal reforming eliminates the cost of an external reformer, and system
efficiency is improved but, as mentioned earlier, at the expense of potentially greater
complexity in cell configuration and issues with catalyst lifetime. Thus there is an
economic compromise or trade‐off to be made between internal and external
reforming.
Internal reforming can only be conducted in an MCFC stack if a steam reforming
catalyst is incorporated. This requirement arises because, although nickel is a good
reforming catalyst, the conventional porous nickel anode has a low surface area and
hence insufficient catalytic activity in itself to support the steam reforming reaction at
the operating temperature (650°C). As will be discussed in Chapter 9, this is not the
case in the SOFC, in which complete internal reforming may be carried out directly on
the anode. For the DIR‐MCFC, the reforming catalyst needs to be close to the anode
to enable the reaction to occur at a sufficiently high rate. Several research groups
demonstrated internal reforming in the MCFC during the 1960s and identified the
major problem areas to be associated with catalyst degradation, which was caused by
carbon deposition, sintering and catalyst poisoning by alkali from the electrolyte.
Internal reforming was studied extensively in the 1990s by BG Technology in a
European Union‐supported programme led by BCN (Dutch Fuel Cell Corporation).
This project identified novel catalyst compositions that could tolerate the presence of
carbonate from the MCFC electrolyte. Key requirements for MCFC reforming catalysts
are as follows:
●
●
●
Sustained activity to achieve the desired cell performance and lifetime. For the
catalyst to provide sufficient activity over the desired lifetime of the stack, any catalyst
degradation must be less than the degradation in the electrochemical performance
of the cell. The reforming reaction is strongly endothermic and thus causes a pronounced
dip in the temperature profile of a cell with internal reforming. This behaviour is
exceptionally severe with the DIR version. Consequently, optimization of the activity
of the reforming catalyst is important to ensure that such temperature variations are
kept to a minimum so as to reduce thermal stress and thus contribute towards a long
stack life. Improvements in temperature distribution across the stack may also be
achieved through the recycling of either the anode gas or the cathode gas, or both.
Resistance to poisons in the fuel. Most raw hydrocarbon fuels that may be used in
MCFC systems (including natural gas) contain impurities (e.g., sulfur compounds)
that are harmful to both the anode and the reforming catalyst. In particular, the
tolerance of most of the catalysts to sulfur is very low, i.e., typically in the parts per
billion (ppb) range.
Resistance to alkali or carbonate poisoning. With DIR, in which the catalyst is located
close to the anode, there is a risk of catalyst degradation through reaction with
carbonate or alkali from the electrolyte. Supported nickel catalysts are generally
preferred, although supported ruthenium has also undergone testing for DIR‐MCFC
application. Poisoning of DIR‐MCFC reforming catalysts is now known to occur
through contact with liquid molten carbonate that arrives via two principal routes: (i)
creepage along the metallic cell components and (ii) transport in the gas phase in the
form of alkali hydroxyl species. The problem is illustrated schematically in Figure 8.7,
which also indicates one possible remedial action, namely, the insertion of a protective
porous shield between the anode and the catalyst.
219
220
Fuel Cell Systems Explained
Electrolyte
Anode
(containing molten
Li2CO3/K2CO3)
Metal cell
housing
Gas phase
transport
LiOH(g)
KOH(g)
Li2CO3(I)
K2CO3(I)
Liquid-phase
creep
Protective
shield
Catalyst
pellets
Figure 8.7 Alkali transport mechanisms in the DIR‐MCFC.
8.4
Performance
The operating conditions for an MCFC are selected essentially on the same basis as
those for a PAFC. The stack size, efficiency, voltage level, load requirement and cost are
all important, and a trade‐off between these factors is usually sought. The performance
curve (voltage vs. current density) is defined by gas composition and utilization, cell
pressure and temperature. State‐of‐the‐art MCFCs generally operate in the range of
100–200 mA cm−2 at 750–900 mV per cell.
As with the PAFC, there is a significant overpotential at the cathode in the MCFC.
This is noticeable if the cell performance when using air as oxidant is compared with
that when using pure oxygen. The resulting behaviour is presented in Figure 8.8 that
shows a cathode performance curve obtained at 650°C with an oxidant that comprised
of oxygen, nitrogen and CO2 as typically used in MCFCs,6 and a curve obtained using a
baseline nitrogen‐free composition. The baseline composition contains the reactants
O2 and CO2 in the stoichiometric ratio that is required for the electrochemical reaction at
the cathode (see Figure 8.1). With this gas composition, little or no diffusion limitations
occur in the cathode because the reactants are provided primarily by bulk flow. The
other (more realistic) gas composition yields a cathode performance that is limited by
gas diffusion and by the lower partial pressure of oxygen in the mixture.
8.4.1
Influence of Pressure
There is a performance improvement to be made by increasing the operating pressure
of the MCFC. As shown in Section 2.5.4, Chapter 2, for a change in system pressure
from P1 to P2, the change in reversible voltage according to the Nernst equation is
given by:
V
RT
P
ln 2
4F
P1
(8.6)
6 The gas composition is the result of burning anode exhaust gas with fresh air. This is the normal means of
supplying CO2 to the cathode inlet of the MCFC. It yields a mixture with a lower oxygen‐to‐nitrogen ratio
than fresh air.
Molten Carbonate Fuel Cells
Baseline cathode gas inlet composition:
33 vol.% O2 + 67 vol.% N2
Cathode overpotential/V
0
–100
Typical MCFC cathode gas inlet composition:
12.6 vol.% O2 + 18.4 vol.% CO2 + 69 vol.%N2
–200
0
100
200
300
400
Current density/mA cm–2
Figure 8.8 Influence of oxidant gas composition on cathode overpotential in a MCFC at 650°C.
(Source: Adapted from Bregoli, LJ and Kunz, HR, 1982, The effect of thickness on molten carbonate fuel
cell cathodes, Journal of the Electrochemical Society, vol. 129(12), pp. 2711–2715.)
From this relationship, it can be shown that, at 650°C, a fivefold and a tenfold increase
in pressure should yield a gain in the open‐circuit voltage of 32 and 46 mV, respectively.
In practice, the increase is somewhat greater because of a reduction in cathode
overpotential. An increase in the operating pressure of MCFCs results in enhanced cell
voltages because of the accompanying increases in the partial pressure of the reactants,
the gas solubilities and the mass‐transport rates.
As was shown when considering the PEMFC at pressure (see Section 4.7.2, Chapter 4),
parasitic power is required to compress reactant gases. Also opposing the benefits of
increased pressure are the effects on undesirable side-reactions such as carbon deposition
(via the Boudouard reaction, as discussed in Section 10.4.4, Chapter 10). Furthermore,
higher pressure inhibits the steam reforming reaction, i.e., equation (7.1), Chapter 7,
which is a disadvantage if internal reforming is being used. These effects, as will be
described in Chapter 10, can be minimized by increasing the steam content of the
fuel stream. In practice, the benefits of pressurized operation are significant only up to
about 0.5 MPa.
The problem of ‘differential pressure’ is another factor to consider. To reduce the risk
of gas crossover between the anode and the cathode in the MCFC, the difference in
pressure between the two sides of each cell should be kept as low as possible. For safety
reasons, the cathode is usually maintained at a slightly higher pressure than the anode
(a few kPa). The ceramic matrix that constrains the electrolyte is a fragile material that
is susceptible to cracking if subjected to stresses induced either through thermal cycling,
temperature variations or excessive pressure differences between the anode and the
cathode. The pressure difference between the anode and the cathode compartments in
stacks has always been a concern of system designers, since recycling of anode burn‐off
gas to the cathode is normally required. Inevitably, there is pressure loss associated with
such gas recycling. The control of small differences in pressure has also militated against
running the stacks at elevated pressures even though there may be advantages from an
efficiency standpoint.
221
222
Fuel Cell Systems Explained
It is also necessary to minimize the pressure difference between the cell compartments and the outside of the stack when running at atmospheric or elevated pressure
since the molten carbonate itself provides the gas‐tight wet seal between the compartments and the outside. Consequently, if an MCFC stack is to operate at elevated
pressure, it must be enclosed within a pressure vessel that is filled a non‐reactive
pressurizing gas, which is usually nitrogen.
Another issue relating to the choice of pressure level is that an improvement in the
overall efficiency of high‐temperature fuel‐cell systems may be achieved through
combination with gas turbines. The latter require hot gas at typically 500 kPa. Solid
oxide fuel cells are very suitable for this application, as they can run in a pressurized
mode and have a high exhaust gas temperature. Molten carbonate fuel cells could also
be combined with gas turbines even though the stack exhaust temperature is lower. For
the reason described earlier, however, the MCFC is not so amenable to operating at high
pressure. Accordingly, although some conceptual systems were devised in the 1990s,
MCFC–turbine facilities are unlikely to be developed.
8.4.2
Influence of Temperature
Simple thermodynamic calculations indicate that the reversible voltage of an MCFC
should decrease with increasing temperature. This relationship is a function of the
change in Gibbs free energy (see Sections 2.1 and 2.2, Chapter 2) and the change in gas
composition at the anode. The main reason for the latter influence is that the gas
composition depends on the equilibrium of the shift reaction, i.e., equation (7.3),
Chapter 7, and this equilibrium is rapidly achieved. The equilibrium constant (Keq) for
the shift reaction increases with temperature; the gas composition therefore changes
with temperature and utilization to affect the cell voltage as illustrated in Table 8.3.
Under actual cell operating conditions, the influence of temperature is most often
dominated by the cathode overpotential. As the temperature is increased, this
overpotential is reduced considerably. The net effect is that the operating voltage of the
Table 8.3 Equilibrium constant (Kc) and equilibrium gas compositions for fuel gas and reversible cell
voltage (Vr) calculated using the Nernst equation for an initial anode gas composition of 77.5 vol.% H2,
19.4 vol.% CO2 and 3.1 vol.% H2O at 0.1 MPa and a cathode composition of 30 vol.% O2, 60 vol. % CO2
and 10 vol.% N2.
Temperature (K)
Parameter
800
900
1000
PH2
0.669
0.649
0.641
PCO2
0.088
0.068
0.052
PCO
0.106
0.126
0.138
PH2 O
0.137
0.157
0.168
Vr (V)
1.155
1.143
1.133
Kc
0.247
0.48
0.711
Molten Carbonate Fuel Cells
MCFC usually increases with temperature.
Above 650°C, however, this effect is very
slight, i.e., only about 0.25 mV per °C. Since
higher temperatures also increase the rate
of all the undesired processes, particularly
electrolyte evaporation and material
corrosion, 650°C is generally regarded as
an optimum operating temperature.
8.5
Practical Systems
8.5.1
Fuel Cell Energy (USA)
In the final years of the 20th century,
MCFC development in the United States
was conducted by two companies, namely,
MC Power and FCE. An earlier programme
undertaken by United Technologies
Corporation (UTC) was concluded in
1992, and, with the consent of the US
Department of Energy, know‐how was
transferred to the Italian company Ansaldo.
The MC Power effort, which grew out of
work performed in the 1960s by the Gas
Technology Institute of Chicago, finished
in 2000. Consequently, FCE became the
sole US manufacturer of MCFC systems.
Figure 8.9 Fuel‐cell stack manufactured by
The company has its headquarters in
Fuel Cell Energy. (Source: Reproduced with
Danbury (CT) and operates a manufacturpermission of Fuel Cell Energy.)
ing plant in Torrington (CT) with a capacity
of 90 MW per year.
All of the products manufactured by FCE incorporate MCFC stacks that typically
comprise 300–400 cells; an example is shown in Figure 8.9. One MCFC stack is used in
the compact DFC300 system (see Figure 8.10) that measures 6 × 4.5 × 6 m3, weighs
19 tons and generates up to 300 kW at 480 V. Exhaust gas flow rate is 1800 kg h−1 at
about 370°C. The system offers a cogeneration capacity between 140 and 235 kW.
The DFC1500 plant (Figure 8.11) produces nominally 1.4 MW of power, measures
16 × 12 × 6 m3 and is built around a fuel‐cell module that houses four stacks. The plant
also features other process modules such as a water treatment skid, a main process skid,
electric balance‐of‐plant and a fuel pretreatment and desulfurization unit. Up to
1100 kW of heat is available from the exhaust gas flow of 8300 kg h−1. The largest of the
FCE products, the 2.8‐MW DFC3000 system (Figure 8.12), is built around two 4‐stack
modules.
In 2014, FCE had installed over 80 sub‐MW and MW‐class DFC power plants around
the world. The facilities have operated successfully on a variety of fuels, such as natural
gas, biogas (digester gas) derived from industrial and/or municipal wastewater, propane
®
®
®
®
223
224
Fuel Cell Systems Explained
Figure 8.10 A 300‐kW DFC‐300® system. (Source: Reproduced with permission of Fuel Cell Energy.)
Exhaust
stack
Water
treatment
and control
panel
Blower,
heater,
preconverter
Fuel
treatment desulfurizer
Module
Switchgear
Inverter
Figure 8.11 DFC1500® 1.4‐MW system showing the different process units. (Source: Reproduced with
permission of Fuel Cell Energy.)
Molten Carbonate Fuel Cells
Figure 8.12 DFC3000® 2.8‐MW system. (Source: Reproduced with permission of Fuel Cell Energy.)
and coal gas. The ‘coal gas’ fuel here includes gas from active and abandoned coal mines
as well as synthesis gas processed from coal.
Biogas offers a unique opportunity for the MCFC. In a wastewater treatment facility,
for instance, the methane‐rich biogas produced by the anaerobic digestion of sludge
becomes the fuel to generate electricity to power the plant, and the resulting exhaust gas
from the fuel cell can be used to heat the sludge to accelerate the anaerobic digestion.
Moreover, biogas is a renewable fuel that is eligible for incentive funding for various
project installations throughout the world. In 2012, FCE had field‐tested several units
on biogas, of which 70% were wastewater treatment applications, the largest being a
1‐MW DFC1500 unit at King County (WA, USA). Unlike natural gas, which is very
consistent in quality, the composition of anaerobic biogas is influenced by the chemical
composition and treatment of the sludge. To achieve a consistent feed required for
stable operation of MCFC stacks, FCE has designed systems in which the biogas is
automatically blended with natural gas.
®
8.5.2
Fuel Cell Energy Solutions (Europe)
Work carried out in the United States during the 1970s and early 1980s by UTC and the
Gas Technology Institute provided the stimulus for European researchers to start their
own research and demonstration programmes in the mid‐1980s. Over the following
two decades, several research and industry groups were formed and delivered world‐
class innovations, but there was reluctance on the part of industry to provide investment
for scale‐up and commercialization. Over a period of 20 years, a body of research was
conducted separately by the Energy Research Centre of the Netherlands (ECN) and
225
226
Fuel Cell Systems Explained
Ansaldo (Italy). Supported by the European Commission under the Framework
Programmes for Research and Technological Development, these organizations collaborated with others including Gaz de France, Sydkraft (Sweden), the Consiglio
Nazionale delle Ricerche (CNR) (Italy), British Gas (BG) Technology (UK), Stork
Engineering (the Netherlands) and the Royal Dutch Schelde Group. Later, the German
company MTU Friedrichshafen started its own MCFC programme in collaboration
with FCE in the United States.
Unfortunately, the European MCFC effort was eventually wound down as follows:
●
●
●
In 2005, ECN sold its interests and intellectual property in MCFC technology to FCE
in the United States.
MTU Friedrichshafen rebranded as CFC Solutions. The prime interest of CFC
Solutions lay in repackaging MCFC technology from the United States for the
European market. The company went out of business in 2010.
Ansaldo work also ended in 2010 with test facilities and intellectual property disbursed
among university research groups in Italy and Poland.
There has, however, been one positive outcome. In May 2012, FuelCell Energy
Solutions GmbH (FCES) was formed as a joint venture between FCE (75%) and the
Fraunhofer Institute for Ceramic Technologies and Systems (Fraunhofer IKTS) (25%).
The venture continues efforts to enhance MCFC technology by combining the strength
of the Direct FuelCell technology developed by FCE with that of the ‘EuroCell’ technology
that will be licensed to the company by Fraunhofer IKTS. The latter builds on the
patents, assets and intellectual property retrieved from the former CFC Solutions and
now held by Fraunhofer IKTS. The assets include the 250‐kW Hot Modules constructed
by MTU and installed throughout Europe. The European venture therefore provides a
platform for FCE in Europe, adding value by accessing former German MCFC research
and development and knowledge of ceramic materials and processing. FuelCell Energy
Solutions GmbH has manufacturing facilities for new European MCFC systems in
Ottobrunn, Germany.
Unlike the FCE stack (shown in Figure 8.9) in which the cells are horizontal, the
European Hot Module is based on the original design by MTU (see Section 8.4.1) and
therefore uses stacks in which the cells are arranged vertically. The features of the MTU
stack arrangement and the Hot Module counterpart, illustrated in Figure 8.13, are as
follows:
●
●
●
●
●
●
●
All hot components are integrated into one common vessel.
Common thermal insulation with internal air-recycling for best temperature
levelling.
Minimum flow resistance and pressure differences.
Horizontal fuel‐cell block with internal reforming.
Gravity‐sealed fuel manifold.
Simple and elegant mechanics.
Transportable by standard truck in sizes up to 400 kW.
The design enables all the processes that need to be run at elevated temperature to be
located within the Hot Module, whereas auxiliaries such as power‐conditioning and
natural gas compression are located outside. In the Hot Module system, a 292‐cell stack
produces about 280 kW DC, which translates to 250 kW AC when power conversion
(a)
Endplate
Insulation
Gas manifold
Fuel
Bipolar plate
Anode
Matrix
Cathode
Oxidator gas
Tension rod
(b)
External
exaust recycle
Hot
module
Cat burner
Mixing zone
Dosing pump
Preheater
FC
Nitrogen
for start and
emergency
DFC
DC/AC convert
grid connector
Fresh air
Nonretum
flap
Startheater
Heat utilization
(optional)
Steam
Boiler
feed water
Hot module
periphery
Fuel gas
Desulfurizer
Electrical Preconverter
startheater
Control system
(c)
Figure 8.13 Example of MCFC system assembly. MTU ‘Hot Module’ cogeneration system showing
(a) stack construction, (b) simplified flow diagram and (c) early demonstration units under
construction. (Source: Reproduced with permission of Fuel Cell Energy.)
228
Fuel Cell Systems Explained
and parasitic losses are taken into account. Exhaust heat is available at about 450°C and
an LHV efficiency of 49% is reported for systems that are fuelled by natural gas.
Two 250‐kW ‘Eurocell’ systems (denoted as DFC 250 EU Reference Plant) have been
deployed at the Elektrizitätswerk Zürich and the German Federal Ministry for Research.
A further two systems (DFC 300 EU Reference Plant) are being installed in London at
two new prestigious developments in Regent Street and Fenchurch Street.
®
®
8.5.3
Facilities in Japan
A MCFC development programme was conducted in Japan from 1981 to 2004. It was a
major undertaking, almost equal to that of the United States. Funding was provided by
Japan’s Ministry of Economy, Trade and Industry (METI) with a total budget of about
US$470 million. The programme was managed by the New Energy Development
Organization (NEDO) and several companies took part, including Fuji Electric, Hitachi,
Ishikawajima‐Harima Heavy Industries (IHI), Mitsubishi Electric, Sanyo Electric and
Toshiba. Many of the participants sought partnership with MCFC developers in the United
States to gain access to expertise. Mitsubishi Electric, for example, formed an alliance
with FCE, and IHI collaborated with MC Power. The Japanese MCFC programme was
dubbed the ‘Moonlight Project’, and there were the following three phases:
Phase 1: (1981–1986) focused on developing 10‐kW stack demonstrations.
Phase 2: (1987–1999) aimed to develop cells and stacks of up to 200 kW, and balance‐of‐
plant systems.
Phase 3: (2000–2004) aimed to develop of a series of high‐pressure, high‐efficiency
short stacks to ultimately develop a 750‐kW system.
A notable success of the Japanese effort was the operation of pressurized stacks at
current densities of 200 mA cm−2 with voltage degradation rates of less than 0.3% per
1000 h, which encouraged the prospect of lifetimes in excess of 40 000 h. Unfortunately,
the programme — particularly the last phase — suffered from various technical setbacks
among the interlinked projects. By the end of the NEDO work in 2004, a projected
7‐MW plant was abandoned as, with the exception of IHI, all of the partners had
withdrawn from the programme. As the sole remaining Japanese company engaged in
MCFC development, IHI started to market MCFC products in 2005, but few were sold.7
It appears that IHI is also no longer engaged in the MCFC business.
8.5.4
Facilities in South Korea
POSCO Energy is the largest private power‐generating company in Korea, with over
40 years of experience in building and operating power plants. The company has
promoted the advancement of MCFCs since early 2000 with government support and
in collaboration with the Korea Electric Power Corporation (KEPCO). Work involved
the development of external‐reforming MCFC technology that in 2010 resulted in the
7 There have been a total of 23 MCFC pre‐commercial prototypes or commercial power plants installed in
Japan. As of April 2012, IHI had four 300‐kW plants in service. At the same time, there were 13 FCE plants
installed, all based on 250‐kW systems. For more details, refer to Fuel Cells by Nobura Behling, listed at the
end of this chapter.
Molten Carbonate Fuel Cells
Figure 8.14 World’s largest operating MCFC power plant (59 MW) located in Hwaseong, South Korea,
comprises 21 DFC‐3000® systems and is owned by Gyeonggi Green Energy. (Source: Reproduced with
permission of Fuel Cell Energy.)
successful demonstration of a 125‐kW plant. In 2007, POSCO Energy was granted a
licence to manufacture, and distribute within Korea, FCE systems that used stacks
supplied from the United States. The aim of the arrangement was to reduce costs by
producing balance-of-plant at a new facility in Pohang. A research and development
centre was established at this site, and in March 2011 a cell manufacturing facility was
added. Current expansion is expected to increase annual output to 170 MW in 2016.
POSCO Power has provided the provinces of Gyeonggi, Geonra, Gyeongsang and
Chungchung with 8.8‐MW MCFCs, and, in 2011, 14‐MW counterparts were supplied to
Suncheon, Dangjin, Ilsan, and Incheon. A 11.2‐MW plant in Daegu City followed in 2012,
and the world’s largest operating fuel‐cell plant was commissioned in 2013 — a 59‐MW
MCFC (see Figure 8.14) that delivers power and district heating to the city of Hwaseong.
As of late 2014, a total of 144.6 MW is being generated by MCFC plants at 18 sites in Korea.
8.6
Future Research and Development
It is clear that the MCFC products now marketed by FCE and associated companies
fulfil a number of commercial requirements. Of all the types of fuel‐cell system, MCFC
power plants are the largest and with over 80 DFC installations worldwide that, to date,
have produced over 2 billion kWh of power. The DFC products are designed to give a
20‐year lifespan with stack lifetimes of between 5 and 7 years. Systems now demonstrate
50% (LHV) efficiencies, low noise (below 65 dBA) and very low emissions (negligible
®
®
229
230
Fuel Cell Systems Explained
with regard to sulfur oxides, SOx, and particulates and less than 10 ppm of nitrogen
oxides, NOx). Nevertheless, with all the advantages that these systems have over alternative technologies, it is still difficult to make a compelling case for the commercial
adoption of MCFC power plant.
It became clear at the close of the 20th century that the capital cost of MCFC systems,
although lower than some other fuel‐cell types, was not low enough for the technology to
be commercially competitive. Investment in scale‐up and manufacturing was required,
and venture capital and other funding sources were difficult to secure. For these reasons,
the programmes in Europe were halted, even though good technical progress had been
made during the previous 20 or so years. A similar situation eventuated in Japan. If MCFC
systems are to become commercially viable in significant numbers, improvements must be
made both in the materials from which cells and stacks are constructed and in the design
of systems. A fundamental issue is the interdisciplinary nature of fuel‐cell research, as will
be discussed in the closing Chapter 12 of this book. It is clear, however, that some areas of
MCFC technology would benefit from further research and investigation, as follows:
●
●
●
Power density. Compared with all other candidate fuel cells, the power density of a
state‐of‐the‐art MCFC is very low. For an internally reforming cell, the power density
is typically 0.16 W cm−2, but with pressurized stacks 0.5 W cm−2 should not be an
unreasonable target.
Cathode overpotential. As with low‐temperature fuel cells, the rate of oxygen reduction
is low and thus leads to poor cell performance. Alternative cathode materials should
be investigated, including metals, oxides and semiconductors.
Wetting of materials by the molten carbonate electrolyte. The corrosion of metals and
alloys by molten salts reduces cell lifetime. A better understanding of passive layer
formation and dissolution–deposition mechanisms is required, especially with
respect to porous or layered composite substrates.
The present configurations of FCE systems are the product of progressive enhancement
in the performance of component materials but with the essential features of internal
reforming that were outlined by Baker and colleagues in the 1960s. As mentioned earlier,
during the late 1990s, the ECN brought in other industrial partners to develop an advanced
DIR‐MCFC system for the European cogeneration market. The consortium devised several
concepts that included the novel method of stack connection illustrated in Figure 8.15. In
this arrangement, three stacks are connected together in series on the cathode side and in
parallel on the anode side. The anode gas is recycled. Calculations show that the system
eliminates the need for all major heat-exchangers and provides a high efficiency. The
concept is also applicable to systems that are comprised of two or more stacks.
Research on various aspects of MCFC technology is being conducted by some universities
and institutes in Europe (Italy, Sweden, Poland, Germany and France), the United States (the
University of Connecticut and Illinois Institute of Technology) and Korea (KIST).
8.7
Hydrogen Production and Carbon Dioxide Separation
For an MCFC system that incorporates the reforming of natural gas or biogas, hydrogen
produced at the anode side is normally consumed by the cell to produce electricity. If
the supply of fuel is maintained but the electrical demand on the system is reduced,
Molten Carbonate Fuel Cells
Recirculation
pump
CH4
Fuel in
Anode
Anode
Anode
Electrolyte
Electrolyte
Electrolyte
Cathode
Cathode
Cathode
Combustor
Exhaust
Air in
Figure 8.15 Stack networking can simplify system design and increase overall performance.
proportionately more hydrogen appears in the anode exhaust gas. Consumption of this
extra gas in the normal cathode burner would increase the temperature at the inlet and
thereby intensify the demand for air cooling. A potentially more useful option is to
separate the hydrogen from the anode exhaust gas so that a change in the cathode air
flow is not required, and the high energy efficiency of the system is maintained. This is
the basis on which FCE has been promoting the DFC systems as hydrogen generators.
Put simply the MCFC can either act as a power generator that is operating at full electrical
load or if the load demand falls, excess hydrogen that would otherwise be routed to the
burner is separated and stored for alternative use (e.g., for fuelling fuel‐cell vehicles).
The concept is attractive for applications where the MCFC system is run in parallel with
an intermittent renewable energy source, e.g., a wind farm.
In the MCFC, there is a transfer of CO2 from the cathode to the anode in the form of
carbonate ions. This transfer can also be put to practical use in that the MCFC can be
employed for separating CO2 from power‐plant flue gases. The idea is also being
promoted by FCE and relies on the fact that the exhaust gas from the cathode of typical
MCFC stacks contains about 1 vol.% CO2 compared with about 10 vol.% at the cathode
inlet. If the cathode is therefore supplied with flue gas from a fossil‐fuel power plant,
most of the CO2 is extracted by the fuel cell and appears in concentrated form at the
outlet of the anode. At the anode outlet, the CO2 may be more amenable to separation
and capture. The concept is illustrated in Figure 8.16.
®
8.8
Direct Carbon Fuel Cell
In 1891, the famous Thomas Edison filed a patent in the United States for the direct
electrochemical conversion of carbon to electricity. The proving of the idea was
contracted to William W. Jacques and Lowell Briggs of the American Bell company, who
in 1896 demonstrated a device for directly producing DC power from baked coal. The
apparatus (Figure 8.17) consisted of 100 cells connected in series and placed on top of a
231
232
Fuel Cell Systems Explained
Hydrogen-rich fuel
CO2
separator
CO2
capture
(~90% CO2)
Recycle or sell
Water
CO2 - depleted
flue gas
(~1% CO2)
Direct
fuel-cell
(DFC)
A
n
o
d
e
C
a
t
h
o
d
e
Supplemental
fuel
Fossil
fuel
Power
Flue gas
(~10% CO2)
Power plant
or process emitting
CO2 greenhouse gas
Power
Air
Figure 8.16 Electricity production and carbon separation. (Source: Reproduced with permission of
Fuel Cell Energy.).)
(a)
(b)
Figure 8.17 Illustration of the Jacques and Briggs carbon battery: (a) 100 cells sitting on top of a coal‐
fired furnace. (b) Details of an individual cell: a carbon, C, is plunged into a solution of caustic soda, E; a
pump, A, forces air into a perforated nozzle, R, which distributes the air uniformly in the electrolyte.
The positive pole is fixed upon an iron receiver, I, that contains the solution, and the negative pole, B,
upon the carbon that is supported and insulated from the receiver by a collar, S. Two tubes, o and i,
serve for the admission and discharge of the electrolyte solution.
Molten Carbonate Fuel Cells
furnace that kept the temperature of the caustic soda electrolyte at between 400 and
500°C. The output was measured as 16 A at 90 V. Since the carbon was consumed rather
than continuously supplied and the hydroxide electrolyte reacted to form carbonate
according to equation (8.7), the system was really a battery rather than a fuel cell.
Current densities as high as 100 mA cm−2 were reported for the earliest cells:
C 2NaOH O2
Na 2 CO3
H2 O
Er
(8.7)
1.42 V
The direct electrochemical oxidation of carbon to CO2 can, in theory, proceed with
very high conversion efficiency since the entropy change of the complete oxidation of
carbon (ΔS°) is very small in comparison with the entropy change for the oxidation of
hydrocarbons:
C O2
G/ H
CO2
1.00;
(8.8)
S
2.5 J mol 1K 1 ; Er
1.0 V ; T
800 C
Since air is the preferred oxidant, the early fused hydroxide electrolytes used in the
Jacques and Briggs cells were later replaced by molten carbonates. Such so‐called direct
carbon fuel cells (DCFCs) employing molten carbonate electrolyte operate at 700–900°C.8
The carbon is oxidized directly to CO2 to yield four electrons per carbon atom. The
half‐cell reactions are as follows:
Anode : C 2CO32
Cathode : O2
2CO2
3CO2
4e
4e
2CO32
(8.9)
(8.10)
Unlike most other types of fuel cell in which both the anode reactants and products
are gases, the cell potential of the DCFC is not dependent on the degree of conversion
of the carbon. Thus, in theory up to 100% carbon conversion could be achieved in
a single‐pass operation. By contrast, if carbon is treated to produce syngas (e.g., by
gasification of coal with steam and oxygen), the syngas converted to hydrogen and the
hydrogen used in fuel cells, then the energy and exergy losses in each step in the process
can add up to a sizeable figure.
After the initial work by Jacques, interest in the carbonate DCFC virtually disappeared
in the first half of the 20th century. It re‐emerged in the late 1990s, prompted mainly by
the high theoretical efficiency of direct carbon oxidation and because of the growing
interest in carbon capture and storage (CCS). Carbon dioxide is the only product at
the anode of the DCFC and thus facilitates its capture, in contrast to conventional coal
gasification or combustion systems that require specific processes for the separation
and capture of CO2.
8 Alternative types of DCFC that have been investigated are based around SOFC materials, that is, with
O2− being the ion that migrates from cathode to anode through a solid yttria‐stabilized zirconia electrolyte.
The operating temperature for this type of cell is in the 800–1000°C range. In such cells the cathode is
lanthanum strontium manganite (LSM), and the anode can be a nickel‐based solid that interacts directly
with fluidized carbon particles. Alternatively the anode can be a molten metal, such as tin, or molten
carbonates into which the carbon fuel is supplied. The latter approach is essentially a hybrid type of molten
carbon/solid oxide fuel cell.
233
234
Fuel Cell Systems Explained
As with the conventional MCFC, CO32− is the ion that migrates from the cathode to
the anode in the carbonate DCFC. In a laboratory‐scale single‐cell DCFC, the ions are
generated by bubbling air through the molten carbonate electrolyte where they interact
with a nickel oxide cathode. It may be more practical in a scaled‐up system to employ a
porous ceramic oxide for the cathode material as in the SOFC, e.g., lanthanum strontium
manganite.
The challenges in developing a practical carbonate DCFC have been the build‐up of
ash in the electrolyte, low anode reaction rates and the high cost of producing suitable
carbon and its transport to the fuel cells. There are also mechanical issues concerning
the distribution of carbon in the fuel cell. The form of carbon has been found to be
particularly influential on the cell performance. It has been shown that carbon with a
highly disordered structure at the nanometre scale (2–30 nm) is much more reactive
than graphite. Such material, commonly referred to as ‘turbostratic’ carbon, can be
obtained by the controlled thermal decomposition of coal, petroleum or natural gas.
The principal challenge for the DCFC is scaling up the concept from a laboratory
batch process to that of a ‘fuel cell’, in which carbon is continually fed and any impurities
such as ash are removed. If the carbonate DCFC is to be developed further, it is also
necessary to show how the structure of carbon is determined by the method of its
preparation and hence how best to produce carbons with high electrochemical reactivity.
Specifically, the key requirement for the DCFC is to devise an efficient and affordable
means whereby a source of carbon, such as coal or biochar, can be transformed into a
low‐ash carbon to serve as a premium fuel.
Further Reading
Behling, N, 2012, History of molten carbonate fuel cells, in Fuel Cells: Current Technology
Challenges and Future Research Needs, pp. 137–221, Elsevier, Amsterdam.
Farooque, M and Maru, H, 2009, Full scale prototypes, in Garche, J, Dyer, CK, Moseley, PT,
Ogumi, Z, Rand, DAJ and Scrosati, B (eds.), Encyclopedia of Electrochemical Power
Sources, vol. 2, pp. 508–518, Elsevier, Amsterdam.
Leto, L, Della Pietra, M, Cigolotti, V and Moreno, A, 2015, International Status of Molten
Carbonate Fuel Cells Technology, Advanced Fuel Cells Implementing Agreement, IEA
Energy Technology Network.
McPhail, SJ, Aarva, A, Devianto, H, Bove, R and Moreno, A, 2011, SOFC and MCFC:
Commonalities and opportunities for integrated research, International Journal of
Hydrogen Energy, vol. 36, pp. 10337–10345.
Selman, JR, 2006, Molten‐salt fuel cells—Technical and economic challenges, Journal of
Power Sources, vol. 160, pp. 852–857.
235
9
Solid Oxide Fuel Cells
9.1
Principles of Operation
9.1.1
High‐Temperature (HT) Cells
The solid oxide fuel cell (SOFC) is a completely solid‐state device that operates with
an oxide ion‐conducting ceramic material as the electrolyte. It is therefore simpler in
concept than all of the other types of fuel‐cell system as only two phases (gas and solid)
are involved. The electrolyte management issues that arise with the phosphoric acid
(PAFC) and molten carbonate (MCFC) fuel cells do not occur, and the high operating
temperatures mean that precious metal electrocatalysts are not necessary. As with the
MCFC, both hydrogen and carbon monoxide (CO) can serve as fuels for the SOFC, as
shown in Figure 9.1.1
The SOFC is similar to the MCFC in that a negatively charged ion (O2−) is transferred
from the cathode through the electrolyte to the anode. Consequently, water is produced
at the anode. Development can be traced back to 1899 when Nernst was the first to
recognize that zirconia (ZrO2) is a conductor of oxygen ions. Until recently, all SOFCs
have been based on an electrolyte of zirconia stabilized with the addition of a small
percentage of yttria (Y2O3). Above a temperature of about 700°C, zirconia becomes a
conductor of oxygen ions (O2−), and state‐of‐the‐art zirconia‐based SOFCs function
between 800 and 1100°C. This is the highest operating temperature range of all fuel cells
and thereby presents extra challenges in terms of construction and durability. Solid
oxide fuel cells typically exhibit good electrical efficiencies, i.e., >50% (LHV), and even
better performance in combined‐cycle schemes. Indeed, both simple‐cycle and hybrid
SOFC plants have demonstrated some of the best efficiencies of any power‐generation
system.
The anode of the SOFC is usually a cermet2 of yttria‐stabilized zirconia (YSZ) and
nickel. The nickel is chosen principally because it has a good electronic conductivity and
is resilient under chemically reducing conditions. Moreover, it is also far more sulfur
resistant than the precious metal catalysts employed in low‐temperature fuel cells and is
1 The high temperatures and presence of steam also means that hydrogen production via the shift reaction
(equation (7.3), Chapter 7) invariably occurs in practical systems, as with the MCFC. The use of the carbon
monoxide may thus be more indirect but just as valuable, as shown in Figure 9.1.
2 A cermet is a composite material that is composed of ceramic (cer) and metallic (met) materials.
Fuel Cell Systems Explained, Third Edition. Andrew L. Dicks and David A. J. Rand.
© 2018 John Wiley & Sons Ltd. Published 2018 by John Wiley & Sons Ltd.
236
Fuel Cell Systems Explained
Product water as steam,
available for steam
reformation of fuel
Hydrogen fuel
Anode
2H2 +
2O2–
→
2H2O +
4e–
O2– ions through electrolyte
Cathode
O2
+
4e–
→
Load
2O2–
Electrons flow round
the external circuit
Oxygen, usually from the air
Carbon monoxide fuel
Anode
2CO
+
2O2–
→
2CO2 +
O2– ions through electrolyte
Cathode
O2
+
4e–
→
4e–
Load
2O2–
Electrons flow round
the external circuit
Oxygen, usually from the air
Figure 9.1 Anode and cathode reactions for the SOFC, when using hydrogen and carbon monoxide fuel.
not poisoned by carbon monoxide. Consequently, the SOFC can accept a wide range
of converted (reformed) fuels, which include coal‐derived gases. Indeed, the presence
of nickel can be used to advantage as a catalyst for internal reforming — it is possible
to perform this process on the anode of the SOFC. By contrast, finding a suitable
material for the cathode has proved to be challenging. In the early days of SOFC
development, noble metals were used but have since fallen out of favour on the grounds
of cost. Most cathodes are now made from electronically conducting oxides or a
ceramic material that possesses both ionic and electronic conductivities. The most
common cathode material of the latter type is strontium‐doped lanthanum manganite
(LSM), La₁₋ₓSrₓMnO₃.
Solid Oxide Fuel Cells
Unlike the MCFC, the SOFC requires no recycling of CO2, which leads to system
simplification, as shown in Figure 9.1. The absence of CO2 at the cathode means
that the open‐circuit voltage (OCV) of the cell is given by the simple form of the
Nernst equation, i.e., equation (2.34), Chapter 2. Nevertheless, the SOFC has one
disadvantage compared with the MCFC in that at the higher operating temperature,
the Gibbs free energy of formation of water is less negative. Consequently, the OCV at
1000°C is about 100 mV less than that of the MCFC at 650°C (see Chapter 2, particularly Figure 2.3 and Table 2.2). The lower OCV would be expected to decrease the
efficiency of the SOFC. In practice, however, the effect is partly offset by the lower
internal resistance of the SOFC and by the ability to use thinner electrolytes than
those required by the MCFC. Consequently, the SOFC can be operated at relatively
high current densities (i.e., up to 1000 mA cm−2).
Development of the SOFC escalated in the 1960s with the introduction of the tubular
design by Westinghouse Electric Corporation. Compared with the planar counterpart,
the tubular shape is more tolerant of the thermal stresses that are associated with the
high temperature of operation. Initially, lanthanum chromite (LaCrO3), a ceramic material,
was used to interconnect the tubular cells. Unfortunately, however, the chromium in
this material can migrate into the cathode and cause serious degradation of the cell.
This behaviour is one of the reasons for the more recent development of intermediate‐
temperature SOFCs (IT‐SOFC).
The high‐temperature SOFC (HT‐SOFC) has been championed by companies such
as Westinghouse Electric, Siemens AG and Rolls‐Royce for application in large‐scale
(base‐load) power‐generation facilities fuelled by natural gas. The ability to carry out
internal reforming potentially simplifies system design and enables an efficiency that is
significantly higher than that obtained by PAFC systems that operate on natural gas.
Furthermore, the heat available from the stack can be used for large‐scale cogeneration
or a combined‐cycle plant.
9.1.2
Low‐Temperature (IT) Cells
Research on HT‐SOFCs during the 1980s and 1990s identified several long‐term
problem areas especially in relation to planar stacks. Delamination of the electrodes from
the electrolyte occurred due to a mismatch between thermal expansion coefficients that
was exacerbated by the thermal stresses encountered at the high temperature of operation.
Sealing was also an issue between adjacent cells and the metal bipolar plates, as well as
between the cells and the metal supporting hardware (see Section 9.2.3.3). Operating
temperatures above 800°C require the use of expensive metal alloys, such as inconel
steels, for the stack hardware and bipolar plates. The problems stimulated researchers
to find ways of reducing the operating temperature of the SOFC. New materials were
identified for the electrolyte and electrodes, and the outcome has been a technology
that operates typically between 600 and 800°C and is now generally known as the ‘IT‐
SOFC’. In pursuing low‐temperature operation, the following advantages were
identified:
●
Metal interconnects, which are subject to severe corrosion at elevated temperatures,
may now be used instead of LaCrO3‐based oxide interconnects; see Section 9.2.3.2.
The use of metals is expected to lead to significant cost reduction and longer
lifetime.
237
238
Fuel Cell Systems Explained
●
●
●
●
●
Thermodynamic conversion efficiency increases for reformed gas (a mixture of
carbon monoxide (CO) and hydrogen).
More options for the sealing of cell components are available.
Low‐cost austenitic steels can be employed in stack construction (i.e., for the bipolar
plates in planar SOFCs) rather than exotic alloys such as inconel metals.
Cell components are less prone to delaminating through differences in thermal
expansion.
For small systems, radiation heat loss from the stack becomes less. Hence, heat
management becomes easier at lower temperatures.
On the other hand, decreasing the operation temperature gives rise to additional
issues with the component materials:
●
●
●
●
The oxide‐ion conductivity of electrolyte material decreases rapidly as the temperature
is reduced. It is therefore essential to have faster oxide‐ion conductors or to develop
a good method of fabricating a thin electrolyte film. Consequently, anode‐supported
electrolytes have proved to be a promising alternative to the more conventional
electrolyte‐supported cell, as discussed in Section 9.3.2.
It is necessary to use electrode materials that have greater activity. For example,
scandium‐doped zirconia is more conductive than YSZ, permitting a further reduction
of the operating temperature of an IT‐SOFC by 50–100°C.
Sulfur poisoning of nickel, which is still the best choice for the anode, becomes more
severe.
Chromium poisoning of the lanthanum strontium manganite used for cathodes
intensifies, against expectations.
In more recent years, another category of cells that operate below 600°C has been
designated as ‘low‐temperature SOFCs’ (LT‐SOFCs).3 Affording the prospect of
faster start‐up and more robust operation compared with high‐temperature
counterparts, these will not be discussed further here as they are at an early stage
of research.
9.2
Components
9.2.1
Zirconia Electrolyte for HT‐Cells
In an SOFC, the electrolyte is exposed to both oxidizing (air side) and reducing (fuel
side) species at high temperatures. Hence, a successful long‐term SOFC operation
requires the electrolyte to have the following properties:
●
●
Sufficient ionic conductivity to minimize ohmic loss and with little electronic
conductivity.
A dense structure, i.e., impermeable to gas. It is difficult to fabricate dense, thin layers
of electrolyte if the porous anode or the cathode is used as the support.
3 Gao, Z, Mogni, LV, Miller, EC, Railsback, JG and Barnett, SA, 2016, A perspective on low‐temperature
solid oxide fuel cells, Energy and Environmental Science, vol. 9, pp. 1602–1644.
Solid Oxide Fuel Cells
●
●
Chemical stability — the electrolyte is exposed to both the air and the fuel at elevated
temperatures, so it must endure both oxidation and reduction processes.
Mechanical compatibility with both electrodes, i.e., the thermal expansion coefficients,
must match at the interfaces.
Zirconia doped with 8–10 mol.% yttria (YSZ) continues to be the most effective
electrolyte for HT‐SOFCs although several other oxides have been investigated, for
example, Bi2O3, CeO2 and Ta2O5. Pure zirconium dioxide (ZrO2) has a monoclinic
crystal structure and is a poor ionic conductor at room temperature. When heated
above 1173°C, it undergoes a phase transformation from monoclinic to tetragonal and
then, on further heating to 2370°C, changes to a cubic fluorite structure. These phases
are ion conductors. The phase change from monoclinic to tetragonal is accompanied by
an appreciable change in volume (about 9%). To stabilize the cubic structure at lower
temperatures and to increase the concentration of oxygen vacancies (which are required
for the conduction of oxygen ions via vacancy hopping) acceptor4 dopants are
introduced into the cation sublattice. Example dopants are Ca2+ and Y3+, which produce
calcia‐stabilized zirconia (CSZ) and YSZ, respectively, as illustrated in Figure 9.2. Each
dopant stabilizes the cubic fluorite structure and improves the oxygen‐ion conductivity
of the zirconia.
Y3+
Zr4+
O2–
ZrO2
Y2O3
Oxygen
vacancy
YSZ (yttria-stabilized zirconia)
cubic fluorite structure
Figure 9.2 Structure of yttria‐stabilized zirconia.
4 In semiconductor physics, a donor atom is one that, when introduced into a semiconductor crystal,
increases the density of electrons to create a so‐called ‘n‐type’ region. The electron density increases
because the donor atom has more electrons in its outer shell than the metal that it is replacing. Conversely,
an acceptor atom creates a deficiency of electrons (known as holes) in a so‐called ‘p‐type’ region. Both
donors and acceptors enhance the electronic conductivity of the material. By analogy, a positively charged
ion, for example, Y3+, is an acceptor dopant for zirconia because it increases the number of oxygen‐ion
vacancies in the crystal lattice.
239
240
Fuel Cell Systems Explained
Zirconia is extremely stable in both the reducing and oxidizing environments,
which are to be found at the anode and cathode regions, respectively, and is a fast
oxygen‐ion conductor above about 700°C. The ability to conduct O2− ions is due to
the replacement of some Zr4+ ions with Y3+ ions in the fluorite structure. When this
exchange of ions occurs, a number of oxygen‐ion sites become vacant because three
O2− ions replace four O2− ions. Transport of oxygen ions occurs between vacancies
located at tetrahedral sites in the lattice, a process that is now well understood at
both the atomic and the molecular levels. The ionic conductivity of YSZ (0.02 S cm−1
at 800°C and 0.1 S cm−1 at 1000°C) is similar to that of liquid electrolytes. Moreover,
YSZ can be made very thin (25–50 µm) and thereby ensures that the ohmic loss
in the SOFC is comparable with that of other types of fuel cell. A small amount of
alumina may be added to the YSZ to improve its mechanical stability, and tetragonal
phase zirconia has also been used likewise to strengthen the electrolyte structure so
that even thinner layers may be fabricated.
The oxygen‐ion conductivity of zirconia can also be influenced by doping. The
atomic number, ionic radius and concentration of the dopant are all found to influence
the conductivity. Among the materials that have been investigated, scandium has
frequently been promoted as a dopant for YSZ, because the ionic radius of the Sc3+
ion in scandium oxide is 0.87 Å and therefore is close to that of the Zr4+ cation in
cubic zirconia (0.84 Å). Although doping YSZ with a small quantity of Sc2O3 does
improve its ionic conductivity, the long‐term performance of this material does not
match that of un‐doped YSZ. Consequently, scandia is normally ruled out as a dopant
on the grounds of cost.
An alternative approach to enhancing the ionic conductivity and performance of
zirconia is to employ a thin film of the electrolyte. Thicknesses as low as 40 µm can be
obtained by electrochemical vapour deposition (EVD), as well as by tape casting and
other ceramic processing techniques. The EVD process was pioneered by Westinghouse
Electric Corporation to produce thin layers of refractory oxides suitable for the
electrolyte, the anode and the interconnection employed in the tubular SOFC design
(v.i.). Now, however, the technique is only applied for fabrication of the electrolyte for
tubular SOFCs. The metal chloride vapour to form the electrolyte is introduced on one
side of the tube surface and an oxygen–steam mixture on the other side. The gas
environments on both sides of the tube act to form two galvanic couples. The net
result is the formation of a dense and uniform metal oxide layer on the tube. The
deposition rate is controlled by the diffusion rate of ionic species and the concentration
of electric charge carriers.
9.2.2
9.2.2.1
Electrolytes for IT‐Cells
Ceria
Zirconia‐based electrolytes are suitable for SOFCs because they exhibit pure anionic
conductivity. Some materials, such as ceria (CeO2) and bismuth oxide (Bi2O3), also
adopt the same fluorite crystal structure. Both of these oxides have greater oxygen‐ion
conductivities than YSZ, but they are less stable at the low partial pressures of oxygen
that pertain at the anode. This condition gives rise to the formation of defects in the
oxides with concomitant increase in the electronic conductivity that, in turn, lowers the
cell voltage.
Solid Oxide Fuel Cells
The ionic conductivity of cerium oxide (CeO2), also known as ceria, can be increased by
doping with gadolinium and can reach a level comparable with that of YSZ at 600°C.5
The doped ceria is known variously in the literature as gadolinium‐doped ceria (GDC),
gadolinia‐doped ceria and cerium‐gadolinium oxide (CGO). Samarium oxide (Sm2O3),
also known as samaria, can equally be used as a dopant to produce samarium‐doped ceria
(SDC) also referred to as samaria‐doped ceria and as cerium‐samarium oxide (CSO). It is the
similar ionic radii of Gd3+/Sm3+ and Ce4+ that gives rise to the enhanced ionic conductivity
of GDC and SDC materials. The main challenge with employing doped ceria is the
reduction of Ce4+ to Ce3+ with the onset of electronic conduction in reducing conditions at
temperatures greater than about 650°C. Further increase in the reduction of Ce4+ to Ce3+
can also lead to lattice expansion and the development of microcracks in the electrolyte.
For bismuth oxide, certain dopants can improve the chemical stability and enhance
ion conductivity, examples being lanthanum, vanadium, copper and zinc. Vanadium‐
and copper‐doped bismuth oxide, for example, has a conductivity that is greater than
that of doped ceria at temperature above about 600°C, but it becomes less stable as the
temperature is increased.
9.2.2.2
Perovskites
As well as oxides with a cubic fluorite structure, perovskites also have a crystal structure
in which oxygen‐ion transport is favoured. The cubic perovskite structure is represented
by the general formula ABO3, where an A‐site ion on the corners of the lattice is usually
an alkaline earth or a rare‐earth element and the B‐site ions in the centre of the lattice
are 3d, 4d or 5d transition metal elements. Either the A or B cation can be substituted
by introducing other cations of the same or different valency. Perhaps the most notable
of the perovskite structures that have been used as electrolytes is that of lanthanum
gallate, illustrated in Figure 9.3. This material, doped with strontium in the A‐sites and
magnesium in the B‐sites, i.e., LaSrGaMgO3, was first reported in 1994.6
Lanthanum gallate is a superior oxide‐ion electrolyte that exhibits pure ionic
conductivity over a very wide range of oxygen partial pressures (10−20 < pO2 < 1). In the
form of strontium–magnesium‐doped lanthanum gallate (LSGM) at 800°C, it provides a
performance that is comparable with that of YSZ at 1000°C, as shown in Figure 9.4.
Unfortunately, the performance declines sharply at lower temperatures as shown in
Figure 9.5. On the other hand, LSGM is also attractive as an electrolyte because it is
compatible with a variety of active cathode materials; hence excellent electrochemical
performance has been reported. The challenges with employing LSGM are the difficulty in
making single‐phase materials and improving both its chemical and mechanical stabilities.
Gallium is volatile at elevated temperatures, and lanthanum can react with nickel (as used
in the conventional SOFC anode) to produce an ionically insulating LaNiO3 phase.
Electrolytes based on gadolinium‐doped ceria, samarium‐doped ceria, LSGM or
other materials such as bismuth metal vanadium oxide (BIMEVOX, Bi4V2O11) have
enabled researchers to focus their attention on IT‐SOFCs and explore a wide range of
materials and fabrication methods for the electrodes.
5 The ionic conductivity of Ce0.9Gd0.1O1.95 is 0.025 Ω−1 cm−1 at 600°C compared with 0.005 Ω−1 cm−1 for YSZ.
6 Ishihara, T, Matsuda, H and Takat, Y, 1994, Doped LaGaO3 Perovskite type oxide as a new oxide ionic
conductor, Journal of the American Chemical Society, vol. 116, pp. 3801–3803; also Feng, M and
Goodenough, JB, 1994, European Journal of Solid State and Inorganic Chemistry, T31, pp. 663–672.
241
Fuel Cell Systems Explained
(a)
(b)
Figure 9.3 Two representations of the cubic perovskite structure of LaGaO3; (a) La‐centred unit cell;
(b) corner‐shared GaO6 octahedra (Ga in the centre surrounded by 6 atoms) with La centred on 12
coordinate sites. Large red spheres = O2− ions; light green spheres = La3+ ions; small blue spheres = Ga3+.
(Source: Ishihara 1994. Reproduced with permission of American Chemical Society.)
Temperature/°C
800
700
600
500
400
300
Stainless steel
Cr-Fe(Y2O3), inconel-Al2O3
Bipolar
plate
material
La(Ca)CrO3
0
Log σ (S cm–1)
242
1500 µm
R0 = L/σ = 0.15 Ω cm2
Bi2 V
–1
150 µm
0.9 Cu
0.1 O
5.35
–2
Self-supported
electrolytes
15 µm
La
Ce
0.9 Gd
(Z
0.9 S
r0
rO
0.1 O
.
G
1.95
1
2)
a
0.9
0.8 M
(Y
g0
2O
.2 O
3 )0
2.8
5
.1
–3
–4
0.8
1.0
1.2
1.4
1.6
1.8
0.5 µm
Supported
electrolytes
0.15 µm
2.0
1000/Temperature/K–1
Figure 9.4 Specific conductivity versus reciprocal temperature for selected solid oxide electrolytes.
The conductivities of some SOFC electrolyte materials are compared in Figure 9.4. To
ensure that the total internal resistance of the cell (i.e., of the electrolyte and electrodes) is
sufficiently small, the target value for the area‐specific resistance (ASR)7 of the electrolyte
in Figure 9.4 is set at 0.15 Ω cm2. Films of oxide electrolytes can be produced reliably using
cheap, conventional ceramic fabrication routes at thicknesses down to approximately
7 The area‐specific resistance (ASR) of a fuel cell is its resistance normalized by its area, ASR = R × A, and is
measured in Ω cm2. An ASR is more fundamental than a resistance because fuel cells are compared on a
per‐area basis. The ohmic losses can be found by multiplying a current density by the ASR. Good fuel cells
target an ASR of 0.1 Ω cm2.
Solid Oxide Fuel Cells
1200
1100
Cell voltage/mV
1000
900
800
700
800°C
600
500
600°C
650°C
700°C
750°C
400
0
200
600
400
800
–2
Current density/mA cm
Figure 9.5 Typical single‐cell performance of LSGM electrolyte (500 µm thick) over a range of
temperatures.
15 µm. It follows that to achieve the aforementioned target for the ASR, the specific
conductivity of the electrolyte must exceed 10−2 S cm−1. This requirement can be met at
500°C for the Ce0.9Gd0.1O1.95 electrolyte and at 700°C for the (ZrO2)0.9(Y2O3)0.1 electrolyte.
Although the Bi2V0.9Cu0.1O5.35 electrolyte exhibits higher conductivity, it is not stable in
the reducing environment imposed by the fuel in the anode region of an SOFC.
Other materials based on perovskite‐related structures that should be mentioned
include Ln2NiO4+δ (where Ln = lanthanum, neodymium or praseodymium). Although
Ln2NiO4+δ materials are mixed ionic and electronic conductors, there is expectation
that it may be possible to suppress their electronic conductivity and thereby open up a
new range of prospective electrolyte materials.
9.2.2.3
Other Materials
In the past few years, two new classes of oxide‐ion conductors have emerged that may
overtake doped ceria and perovskite materials for IT‐SOFC electrolytes. The first is
lanthanum molybdenum oxide (LAMOX, La2Mo2O9). This material exists in several
phases, and a high‐temperature cubic form possesses exceptionally good oxygen‐ion
conductivity in comparison with most of the other perovskite materials. Apatite‐type
oxides are the other class of alternative electrolyte materials. The general formula of apatite
oxides can be written as M10(XO4)6O2±y where M is a rare‐earth or alkaline earth cation;
X is a p‐block element such as phosphorous, silicon or germanium; and y is the amount
of oxygen non‐stoichiometry. These oxides have the same structure as hydroxyapatite
materials found in bones and teeth. The apatites that incorporate silicon or germanium
are among the most promising of the new fast‐ion conductors, as shown in Figure 9.6.
9.2.3
9.2.3.1
Anodes
Nickel–YSZ
In the nickel–YSZ cermet that forms the anode of state‐of‐the‐art HT‐SOFCs, the
porous YSZ serves to inhibit the coarsening or sintering of the metal particles (that
otherwise would lead to loss of surface area) and provides the anode with a thermal
243
Fuel Cell Systems Explained
–1
Log σ (S cm–1)
244
–2
YSZ
Bi2V0.9Cu0.1O5.35
Cerium-gadolinium oxide
–3
Ge-apatite
Silicon-apatite
–4
LSGM
LAMOX
–5
0.8
1.0
1.2
1.4
1.6
1000/Temperature/K–1
Figure 9.6 Total conductivities of several well‐known oxide‐ion conductors as a function of reciprocal
temperature: CGO, Ce0.8Gd0.2O1.9; Ge‐apatite, La10(GeO4)6O3; LAMOX, La2Mo2O9; LSGM, La0.9Sr0.1Ga0.8
Mg0.2O2.85; Si‐apatite, La10(SiO4)6O3 and YSZ, (ZrO2)0.92(Y2O3)0.08.
expansion coefficient similar to that of the electrolyte. At least 30 vol.% of nickel is
generally used in the cermet in order to achieve adequate electronic conductivity while
maintaining the required porosity so that mass transport of reactant and product gases
is not compromised. Nickel and YSZ are non‐reactive over a wide range of temperature
and are also essentially immiscible in each other. Both these properties simplify the
synthesis of YSZ cermets and allow the preparation through conventional sintering of
NiO and YSZ powders. Reduction of NiO to Ni in situ leads to a highly porous (20–
40%) YSZ structure that contains connected Ni particles on the surface of the YSZ
pores. The structure provides a substantial three‐phase boundary for oxidation of fuels
together with paths for electronic conduction from the reaction zone to the SOFC
current-collector.
Traditionally, the anode is generally prepared by painting an ink made from the
NiO and YSZ powders directly onto the YSZ electrolyte tube or plate, according to
the cell configuration required. The painted electrolyte is dried and then sintered at
approximately 1400°C in air during which the aforementioned porous anode structure
forms as the result of the growth of the NiO and YSZ particles. Sintering at such an
elevated temperature also produces an excellent bond between the resulting porous
anode layer and the dense YSZ electrolyte layer.
Anode‐supported electrolyte cells are a relatively recent innovation for the IT‐SOFC.
In these cells, the anode is prepared by extruding or pressing NiO and YSZ powders
into the required shape, which is then dried sintered to produce a relatively thick plate
that can support a thin layer of electrolyte. Powdered electrolyte material is painted
onto the anode and the two layers are co‐fired as before.
Factors that affect the performance of a Ni–YSZ cermet anode include the properties
of the starting materials, the sintering temperature and the nickel content. In addition,
modification of the microstructure and doping with various materials can improve
long‐term performance by lowering the degree of sintering of nickel particles. To this
Solid Oxide Fuel Cells
end, MgO, TiO2, Mn3O4 and Cr2O3 have all been shown to reduce nickel sintering and
also to act as anchor sites for nickel at the anode|electrolyte interface.
As noted previously, hydrocarbon fuels can be reformed directly on the SOFC anode.
While this feature carries advantages in terms of greater system efficiency, there are
risks involved: notably, the formation of carbon either by the Boudouard reaction or
through cracking (pyrolysis) on the nickel (see Section 10.4.4, Chapter 10). Investigations
have shown that the propensity for carbon formation on an SOFC anode can be reduced
by doping the cermet with molybdenum or gold.
There is some ohmic loss at the interface between the anode and the electrolyte, and
several developers have investigated bilayer anodes in which a small amount of ceria is
added to the Ni–YSZ cermet. This improves the tolerance of the anode to temperature
variations and to cycling between oxidation and reducing conditions on the surface
(changing the anode gas from a reducing fuel gas to an oxidizing gas and vice versa).
Control of the particle size of the YSZ can also improve the stability of the anode
towards oxidation and reducing conditions.
9.2.3.2
Cathode
Ceria itself possesses a substantial surface concentration of O2− ions that renders it a
good catalyst for hydrocarbon oxidation. By doping ceria with certain rare‐earth
metals, the catalytic activity for oxidation is enhanced. In addition, the doped ceria has
a fluorite structure with not only conductivity for oxygen ions but also some electronic
conductivity. Consequently, ceria has been pursued as an anode material for SOFCs in
its own right. Although doped ceria has a good ionic conductivity, the electronic
conductivity is relatively low, and therefore the material has most commonly been
employed in the form of a cermet. The combination of doped ceria and metal provides
excellent ionic conductivity at low temperature and a high electronic conductivity, and
therefore the material is an ideal candidate for IT‐SOFC applications.
Gadolinia‐doped ceria combined with nickel to form a Ni–GDC cermet has been
shown to have high electrochemical activity towards the steam reforming of methane at
temperatures as low as 500°C without coking of the catalyst. The principal problem with
the material is the possibility of mechanical degradation through expansion of the crystal
lattice, which results from the bulk transition of Ce4+ to Ce3+ that may occur in a low
oxygen partial‐pressure environment. This is of particular concern when a YSZ electrolyte
is used, as delamination at the electrolyte|electrode interface could occur. Doping the
ceria with cations of low valency such as Sm3+, Y3+ or Gd3+ may help to prevent the
degradation, and doping a ceria anode with 40–50 at.% of gadolinia, e.g., Ce0.6Gd0.4O1.8,
can provide a good compromise between electronic conductivity and mechanical stability.
Doped‐ceria cermets have been shown to sustain multiple, rapid thermal cycles and
full oxidation–reduction cycles without a decline in performance. They can be employed
in both HT‐SOFCs and IT‐SOFCs. Reaction between the cermet and YSZ electrolyte in
the HT‐SOFC can be prevented by inserting a thin interlayer of ceria doped with both
gadolinia and zirconia between the electrolyte and the ceria cermet anode.
A problem with using cermets that incorporate ceria in conjunction with a ceria
electrolyte is the reduction of Ce4+ to Ce3+, mentioned previously, and the ensuing
increase in electronic conductivity. For this reason some developers insert a thin
interlayer of ion‐conducting oxide material at the junction between the anode and the
electrolyte of an IT‐SOFC. Alternatively the surface of the anode that bonds with the
245
246
Fuel Cell Systems Explained
electrolyte can be functionalized to limit its electronic conductivity. Both approaches
protect the ceria electrolyte from the reducing conditions at the anode.
When using hydrocarbon fuels, the nickel in anode cermets, while providing good
electronic conductivity, suffers the disadvantage of promoting carbon formation. Nickel is
also poisoned by sulfur. Both issues are a concern if internal reforming of hydrocarbons is
undertaken. To eliminate these problems, nickel can be replaced by copper in ceria
cermets doped with rare earths, as described in Section 10.4.6, Chapter 10. A composite
anode consisting of copper, ceria and YSZ has shown activity for direct electrochemical
oxidation of hydrocarbons without any performance degradation by carbon deposition.
When compared with a Ni–YSZ anode, the composite shows superior performance with
both carbon monoxide and hydrocarbon fuels. The material can be improved further by
incorporating noble metals such as platinum, rhodium or palladium.
9.2.3.3
Mixed Ionic–Electronic Conductor Anode
With the development of electrolyte materials for IT‐SOFCs, attention has been given to
alternative anode materials to the conventional cermets. Certain perovskites possess both
electronic and ionic conductivities. Such materials are known as ‘mixed ionic–electronic
conductors’ (MIECs). Examples that have been investigated include titanates such as
strontium titanate (SrTiO3) doped with either yttrium or lanthanum, A‐site deficient
lanthanides doped with neodymium and praseodymium and iron‐doped barium titanate.
Perovskites based on chromites have also been examined, e.g., strontium‐doped lanthanum
chromite doped with transition metals such as Co, Cu, Ni and Mn. Vanadium complexes
and ferrites have shown some promise as anodes, e.g., strontium‐doped lanthanum
vanadate La1–xSrxVO3 (denoted as LSV), as have ferrites such as lanthanum strontium
cobaltite ferrite (LSCF). Needless to say, the search for effective and stable SOFC anodes is
a very active field of research. Apart from the elimination of the metal (e.g., nickel) and
therefore a reduction in the chance of carbon deposition, MIEC materials provide a means of
extending the three‐phase boundary between anode and electrolyte, as shown in Figure 9.7.
(a) Electronically conducting cermet
(b) Mixed ionic–electronic conductor
H2
H2
H2O
H2O
O=
e–
Electrolyte
e–
Particle of a
pure electronic
conductor
Three-phase
boundary
region
O=
O= e–
H+
H+
Electrolyte
e–
Particle of a
mixed ionic
and electronic
conductor
Figure 9.7 (a) Illustration of the three‐phase boundary regions of different SOFC anode materials.
(b) Extension of the boundary is obtained with mixed conducting cathode materials.
Solid Oxide Fuel Cells
9.2.4
Cathode
The very first SOFCs used platinum for the cathode, but as soon as electronically
conducting ceramics became available, these were favourably received. The first to
attract serious attention was the perovskite lanthanum cobaltite (LaCoO3). Initially,
this compound exhibited a good performance, but longer tests showed that it
reacted with the YSZ electrolyte to cause permanent degradation. Although the ionic
conductivity could be improved by doping with strontium, the LaCoO3 remained
reactive towards YSZ. It is non‐reactive with doped‐ceria electrolytes but has a poor
match with the expansion coefficient of YSZ. Other cobaltite materials were
investigated and included gadolinium strontium cobaltite Gd1‐xSrxCoO3 that was also
suitable for use with ceria electrolytes. After exploring the cobaltites for cathode
materials, researchers moved to the investigation of manganites, and LSM emerged
as the preferred material. Despite being susceptible to some reaction with YSZ,
particularly at elevated temperatures, strontium‐doped LSM has become the most
widely used cathode material for the SOFC over the past 20 years for both HT‐ and
IT‐SOFCs. As with many perovskites, the crystal structure of LSM undergoes a phase
change from an orthorhombic structure at room temperature to a rhombohedral
form above 600°C. The mixed ionic–electronic conductivity of LSM can be enhanced
by replacing some of the A‐sites with Sr2+. It is particularly notable that the defect
structure and oxygen non‐stoichiometry of LSM depend on the applied partial pressure
of oxygen. The fact that LSM can have oxygen excess or oxygen deficiency is unusual
in comparison with most perovskite oxides. The material is stable over a wide range
of oxygen partial pressures, but at very low levels, it can decompose to form two
phases, namely, La2O and MnO.
Although LSM has proven itself for most HT‐SOFCs, other materials have also been
found suitable for the cathode, particularly those with perovskite structures that
exhibit mixed ionic and electronic conductivity. Mixed conductivity is especially
important for operation at lower temperatures since the cathode overpotential
increases significantly with reduction of the SOFC temperature. The advantages of
using oxides with mixed conductivity become especially noticeable in cells that
operate at 650°C. Lanthanum strontium ferrite and lanthanum strontium cobaltite,
which are all n‐type semiconductors, are better electrocatalysts than the state‐of‐the‐
art lanthanum strontium manganite at IT‐SOFC temperatures.
9.2.5
Interconnect Material
The interconnect is the means by which neighbouring fuel cells are joined together. In
the planar design of cell, this is the bipolar plate, but the arrangement is different for
tubular geometries, as will be described in Section 9.3.1. Doped lanthanum chromite
has been the material of choice for the interconnect in HT‐SOFCs, notably in the
Westinghouse Electric tubular stacks. The electronic conductivity of pure chromite is
very low but may be enhanced by substituting ions of lower valency (e.g., calcium,
magnesium, strontium) on either the La3+ or the Cr3+ sites of the lanthanum chromite
lattice. Unfortunately, the material has to be sintered at quite high temperatures
(1625°C) to produce a dense phase. This requirement exposes one of the major problem
areas with the SOFC, namely, the means of fabrication. All of the cell components must
247
248
Fuel Cell Systems Explained
be compatible with respect to chemical stability and mechanical compliance (e.g., they
must all have similar thermal expansion coefficients). The various layers have to be
deposited in such a way that good adherence is achieved without degrading the material
through the use of too high a sintering temperature. Many of the methods of fabrication
are proprietary, and considerable research is being conducted on the processing of
SOFC materials.
In cells intended for operation at lower temperatures (<800°C), it is possible to use
oxidation‐resistant metallic alloys for the interconnects. Ferritic steels are currently
the preferred option. Compared with lanthanum chromite ceramic, metallic alloys
offer advantages such as improved manufacturability, significantly lower raw material
and fabrication costs and superior electrical and thermal conductivity. Alloys, such as
the Cr‐5Fe‐Y2O3 Plansee material and Crofer22 APU (a high‐temperature ferritic
stainless steel developed by VDM Metals GmbH), have been engineered with thermal
expansion coefficients to match those of SOFC ceramic components. Unfortunately,
under high oxygen partial pressure at the cathode, the chromium in such alloys has a
tendency to vapourize and deposit at the LSM–YSZ three‐phase boundary and
thereby cause permanent poisoning of the cathode. Another advantage for IT‐SOFCs
is that cheaper low‐nickel materials, such as austenitic steels, may also be used in the
manufacture of cells.
9.2.6
Sealing Materials
A key issue with SOFCs, particularly planar configurations, is the method of sealing
the ceramic and metal components to obtain gas tightness. The wet‐seal arrangement
employed in the MCFC cannot be used, and the wide temperature range of SOFC
operation poses particular problems. The seal needs to be thermochemically stable
and compatible with other cell components; it should also withstand thermal
cycling between room temperature and the operating temperature. To satisfy such
requirements, a number of different sealing approaches have been developed — some
more successful than others — and include rigid bonded seals (e.g., glass ceramics
and brazes), compliant seals (e.g., glasses) and compressive seals (e.g., mica‐based
composites).
The most common practice has been to use glasses that have transition temperatures
close to the operating temperature of the cell. These materials soften as the cells are
heated up and form a seal all around the cell. Glass seals are employed in planar
stack designs in which, for example, a number of cells may be assembled in one
layer (v.i.). A particular problem associated with the use of glass seals is the
migration of silica from the glass, especially onto the anodes with resultant degradation
in cell performance. Glass ceramics have been adopted for all‐ceramic stacks, but
migration of the silica component can still be a problem on both the anode and
cathode sides.
Brazes have not been featured strongly as sealants because of the ease by which
suitable metals can oxidize at the elevated temperatures necessary for carrying out
the brazing operation (i.e., above 800°C). Oxidation can be reduced by adding a
metal such as titanium or zirconium to the brazing metal, but the need for a reducing
atmosphere during the brazing process can reduce the activity of cathode
components.
Solid Oxide Fuel Cells
Compliant compressive seals (i.e., gaskets) have had limited application in SOFCs as
most of the appropriate metals tend to oxidize or deform excessively at the operating
temperature under a sustained compressive force. This is usually less of a problem with
short‐term laboratory tests where, for example, gold gaskets have been routinely used.
Mica composite materials of greater resilience have been employed for single‐cell tests
and may find application in future planar stack configurations.
9.3
Practical Design and Stacking Arrangements
9.3.1 Tubular Design
The tubular SOFC was pioneered in the United States by the Westinghouse Electric
Corporation (now the Siemens Westinghouse Power Corporation). The original design
used a porous CSZ support tube of 1–2 mm thickness and about 20 mm internal
diameter, onto which the cylindrical anodes were deposited. By a process of masking,
the electrolyte, the interconnect and finally the fuel electrode were deposited on top of
the anode. The procedure was reversed in the early 1980s so that the air electrode
became the first layer to be deposited on the zirconia tube, and the fuel electrode was
on the outside. This tubular fuel cell became the standard for the next 15 years. From its
inception, the tubular design has suffered from the major problems such as low‐power
density and punitive fabrication costs. The low‐power density resulted from both the
long path for the electrical power through each cell, as depicted in Figure 9.8, and the
large voids within the stack structure (i.e., between the tubes). The unfavourable costs
arise from the preparation of the electrolyte and electrode via electrostatic vapour
deposition, which is a batch process conducted in a vacuum chamber.
More recently, the zirconia support tube has been replaced by the one made by the
extrusion of a porous form of doped LSM onto which the electrolyte is deposited by
EVD, followed by plasma spraying of the anode. The arrangement is known as an ‘air‐
electrode supported’ (AES) fuel cell; an example is shown in Figure 9.9.
One significant advantage of the tubular design of SOFC is that high‐temperature
gas‐tight seals are eliminated, as illustrated in Figure 9.10. Each fuel‐cell tube is
fabricated in the form of a large test tube, sealed at one end. Fuel flows along the outside
of the tube (anode side) towards the open end. Air is fed through a thin alumina supply
tube that is located centrally inside each tubular fuel cell. Heat generated within the cell
brings the air up to the operating temperature. The air then flows through the fuel cell
back up to the open end. At this point, air and unspent fuel from the anode exhaust are
instantly combusted so that the cell exit is above 1000°C. This combustion provides
additional heat to preheat the air supply tube. Thus, the tubular SOFC has a built‐in air
preheater and anode exhaust gas combustor, as well as no requirement for high‐
temperature seals. Moreover, by allowing imperfect sealing around the fuel‐cell tube,
some recirculation of anode product gas (which contains both steam and CO2) occurs
and thereby allows internal reforming of fuel gas on the SOFC anode.
The current SOFC tubes produced by the Siemens Westinghouse Power Corporation
are 150 cm in length and 2.2 cm in diameter. The voltage versus current density and
power versus current density characteristics of a single‐tube cell at 900, 940 and 1000°C
and operating with 89 vol.% H2 + 11 vol.% H2O fuel (85% fuel utilization) and air as
249
250
Fuel Cell Systems Explained
More cells
Fuel
Fuel
Electrolyte
Air
Typical path of an
electron from the anode
of one cell to the cathode
of the next
Cell interconnect
Fuel
Fuel
Anode
Air
Cathode
More cells
Figure 9.8 End view of tubular SOFC produced by Siemens Westinghouse. Both electrolyte and anode
are built onto the air cathode.
Interconnect
Electrolyte
Cathode
Fuel
flow
Cell bundle
of 3 × 8 tubes
Airflow
Anode
Figure 9.9 Small stack of 24 tubular SOFCs. Each tube has a diameter of 22 mm and is about 150 cm
long. (Photograph reproduced by kind permission of Siemens Westinghouse Power Corporation).
Solid Oxide Fuel Cells
This seal or joint is
quite straightforward
to maintain
Exhaust
Air
Fuel
Combustion
Deliberately
‘imperfect’
seals around
the tube
Recirculation
A
i
r
Tubular
fuel
cell
Fuel
A
i
r
Internally
reformed
using
product
steam
Figure 9.10 Diagram showing how the tubular‐type SOFC can be constructed with (almost) no seals.
oxidant are presented in Figure 9.11. Such tubular cells have a power density at 1000°C
of about 0.2 W cm−2, i.e., much lower than that of planar cells. Individual tubular cells
are arranged in series–parallel stacks of 24 tubes, as displayed in Figure 9.9.
Several other organizations — notably, Mitsubishi Heavy Industries and TOTO Ltd.
in Japan, Adelan Ltd. in the United Kingdom and both Acumentrics and Watt Fuel
Cell in the United States — have been developing tubular SOFC designs. To avoid the
expensive EVD process, TOTO Ltd. has adopted wet sintering as a method of cell
fabrication.
9.3.2
Planar Design
Alternatives to the tubular SOFC have been pursued for several years, especially several
types of planar configuration and a monolithic design. In the planar designs, cells are
thin flat plates that are electrically connected to achieve the required stack voltage and
current. The earliest planar cells featured the electrolyte as the support onto which the
electrodes were deposited. More recently and especially for IT‐SOFCs, the anode‐
supported structure is preferred in which the anode and the electrolyte are deposited
directly onto a metal bipolar plate and above the anode, respectively. There are several
251
Fuel Cell Systems Explained
1.2
800
1.0
600
0.9
0.8
400
0.7
800
750
700
650
600
0.6
0.5
0.4
0
200
400
600
800
Current density/mA cm–2
1000
Power density /mW cm–2
1.1
Voltage/V
252
200
0
1200
Figure 9.11 Influence of temperature on the performance of Siemens Westinghouse tubular fuel cells.
variants of the planar design. In one example, promoted by the Swiss company Sulzer
Hexis AG, the SOFC is in the form of a circular disc fed with fuel from the central axis,
whereas in another design, preferred by Siemens AG and others, the cell employs a
square plate that is fed from the edges.
Planar designs offer several prospective advantages that include simpler and less
expensive manufacturing processes and power densities that are better than those
achievable with tubular SOFCs. Unlike tubular cells, however, planar designs require
high‐temperature gas‐tight seals between the components in the SOFC stack. Sealing
therefore remains one of the most significant technical barriers to the commercialization of planar SOFCs. Also of concern are the thermal stresses at the interfaces between
the different cell and stack materials that can cause mechanical degradation. Particularly
challenging is the brittleness of thin planar SOFCs in tension. Thermal cycling is a
further problem for the planar SOFC; by contrast, the tubular cell is thermally more
robust. Finally, the issue of thermal stresses and the fabrication of very thin components
place a major constraint on the size of planar SOFCs. Early configurations employed a
thick electrolyte as the support, which required an operating temperature that was
often taken above 900°C to achieve adequate current densities.
Advances in ceramic processing have allowed reproducible fabrication of thin
electrolytes, i.e., 10 µm or thinner, by low‐cost conventional ceramic processing
techniques such as tape casting, tape calendaring, slurry sintering and screen printing
or by plasma spraying. The ability to create thin electrolytes led to the interest in anode‐
supported cells. For many years, the maximum size of single planar SOFCs was
5 cm × 5 cm. Now larger cells can be routinely manufactured and assembled into a stack
of much greater area by building them into a window‐frame arrangement such that
several cells are located in one layer. An example arrangement that uses four cells in
each layer is shown in Figure 9.12.
Solid Oxide Fuel Cells
Figure 9.12 Stack with four 10 × 10 cm2 SOFCs in one layer (window‐frame design FY520). (Source:
From Blum, L, Batfalsky, P, Fang, Q, deHaart, LGJ, Malzbender, J, Margaritis, N, Menzler, NH and Peters,
R, 2015, SOFC stack and system development at Forschungszentrum Jülich, Journal of the
Electrochemical Society, vol. 162(10), pp. F1199–F1205.)
9.4
Performance
When hydrogen is the fuel, the OCV of the SOFC is lower than that of the MCFC and
PAFC. Nevertheless, at the operating temperature of the SOFC, the overpotential at the
cathode is much lower and thereby provides the technology with a superior operational
voltage. The voltage losses in SOFCs are a function mainly of the resistance of the cell
components, which include those associated with current collection. The contributions
to ohmic voltage loss in a tubular cell are typically some 45% from the cathode, 18%
from the anode, 12% from the electrolyte and 25% from the interconnect, when these
components have a thickness of 2.2, 0.1, 0.04 and 0.085 mm and resistivities at 1000°C
of 0.013, 3 × 10−6, 10 and 1 Ω cm, respectively. With a tubular SOFC, the ohmic loss at
the cathode dominates despite the greater resistivities of both the electrolyte and the
cell interconnection. This situation arises because the conduction path through these
latter two components is much shorter than the current path in the plane of the cathode,
as demonstrated in Figure 9.5. With planar cells, the long current path is absent, and
therefore it is possible to achieve higher power outputs from the stacks.
As with other types of fuel cell, SOFCs show enhanced performance with increase in
cell pressure. Unlike low‐ and medium‐temperature cells, however, the improvement is
mainly due to the increase in the Nernst voltage. In Section 2.5.4, Chapter 2, it was
shown that the voltage change for an increase in pressure from P1 to P2 follows very
closely the theoretical equation
V
0.027 ln
P2
P1
(2.45)
The relationship was borne out in practice when Siemens Westinghouse in
conjunction with Ontario Hydro Technologies tested AES tubular cells at pressures
of up to 15 atm (1.52 MPa) on both hydrogen and natural gas. Operation at such
pressures is particularly advantageous when using the SOFC in a combined‐cycle
system with a gas turbine. In other cases, as with the proton‐exchange membrane
fuel cell (PEMFC), the power costs involved in compressing the reactants render the
benefits marginal.
253
254
Fuel Cell Systems Explained
The temperature of an SOFC has a very marked influence on its performance, though
the details will vary greatly between cell types and the materials used. The improvement
in the performance of the Westinghouse tubular fuel cells on increasing the temperature
from 900 to 1000°C is demonstrated by the data given in Figure 9.11. The predominant
effect is that high temperatures increase the conductivity of the materials, and this
reduces the ohmic losses within the cell that, as discussed in Section 3.65 Chapter 3, are
the most important type of loss in the SOFC.
For SOFC–combined‐cycle and hybrid systems, it is beneficial to maintain a high
operating temperature. For other applications, such as cogeneration and possible
transport applications (e.g., as an auxiliary power supply for vehicles), it is more
beneficial to operate at lower temperatures, as the higher temperatures bring material
and construction difficulties. Unfortunately, as Figure 9.11 clearly shows, for a given set
of cell components, the performance decreases substantially for SOFCs as the temperature
is lowered.
9.5
Developmental and Commercial Systems
Reference was made in Chapter 8 to the significant MCFC development programmes
in Europe and Japan that have been abandoned despite the investment of substantial
public and private funding over many years. A similar theme has been seen with SOFC
programmes. In 1977, the US Department of Energy commenced its funding for the
development of the Westinghouse Electric tubular fuel cell. Given the issues described
in the previous section (e.g., a long current path that led to low stack power and low
power-density), Westinghouse made good progress throughout the 1980s, and the
company believed that commercialization would be possible by the early 1990s. A
number of set‐backs, which were partly technical and partly due to restructuring of the
company at the turn of the century, slowed down the progress of the technology, but
work continued until 2010. By that time, as noted previously, Siemens had taken over
the fuel‐cell interests of Westinghouse Electric (with the creation of the Siemens
Westinghouse Power Corporation), and, after many reviews of their progress, it was
decided that SOFC development was not the core business for the parent company
(Siemens in Germany). Consequently, further activities were stopped and the fuel‐cell
enterprise was put up for sale. More details of the history of Siemens and various SOFC
developers are catalogued by Behling8.
Continued interest in the technology was assisted by the US Department of Energy,
which in 2001 instigated a programme that focused on planar SOFCs. Known as the
Solid State Energy Conversion Alliance (SECA), the programme was intended to
develop an SOFC system that would cost as little as US$400 per kW by 2010. Whereas
on most counts the SECA programme failed to live up to expectations, it did serve to
restimulate activity in SOFCs, especially for planar systems — both in the United States
and around the world.
The present status of the SOFC is very different from that of the MCFC. In the case
of MCFC development, few players are left, whereas for the SOFC many are active in
8 Behling, N, 2023, History of solid oxide fuel cells, in Fuel Cells, Current Technology Challenges and Future
Research Needs, Elsevier, Amsterdam. ISBN 9780444563255
Solid Oxide Fuel Cells
research and demonstration as well as commercialization, and, moreover, several of the
participants are relatively new to the field. The following selected SOFC systems are
currently under development or being commercialized. The examples are not intended
to be an exhaustive list but to illustrate the wide variety of approaches that companies
have taken. As well as Siemens Westinghouse Power Corporation, some other companies
that appeared to be well advanced have pulled out of SOFC development in recent
years. It therefore should be recognized that the industry is very fragile. The business
landscape could change rapidly over the next few years as some research teams push
forward and others fall from view.
9.5.1 Tubular SOFCs
Despite the exit of Westinghouse from the fuel‐cell scene, there remain several
developers of tubular SOFC. Most activities are focused on small systems for remote
backup power, auxiliary power units for vehicles, and portable power supplies. The
participating companies include Watt Fuel Cell, Acumentrics, Inc. and Protonex
Technology Corporation in the United States and Adelan Ltd. in the United Kingdom.
Acumentrics, Inc. was formed in 1994 to develop rugged uninterruptible power
systems for use in harsh environments, and the company was involved in the SECA
programme until 2010. Acumentrics has one of the most advanced tubular SOFC
technologies that is perhaps closest in form to the Westinghouse tubular cells. The
company has deployed over 350 remote power generators in a variety of sites throughout
North America, and in 2015 the successful fuel‐cell activities were divested into a new
venture — Atrex Energy, Inc. With a portfolio of four SOFC products that have outputs
of 250, 500, 1000 and 1500 W, respectively, Atrex Energy has shipped systems to Europe
and Asia as well as fulfilling a growing demand in North America. The technology
differs from the original Westinghouse tubular cell in that fuel, rather than air, is fed
through the centre of the tube, and therefore, the anode is on the inside of the tube and
the cathode on the outside. Moreover, the anode (Ni–YSZ) extends the full internal
length of the tube, whereas the cathode (LSM) is segmented into discrete partitions
so as to enable efficient current collection from the anode at different stages. This
arrangement avoids electrons having to flow throughout the whole length of the cell
that, otherwise, would lead to greater ohmic losses. LaCrO3 wires coiled around the
electrode partitions act as current-collectors; moreover, a tight coiling around the
anode prevents an undesirable exposure to the air stream that could potentially result
in reoxidation of the nickel matrix and, in turn, cause mechanical stress. A bundle of
Atrex Energy/Acumentrics tubes is shown in Figure 9.13. Solid oxide fuel cells produced by Atrex Energy and others such as Protonex are intended for remote stationary
systems and are fuelled with propane.
Several companies in Japan have ongoing SOFC programmes. Kyocera Corporation
and Mitsubishi Hitachi Power Systems (MHPS) have been in the fuel‐cell field for
several decades and, with support from central government funds, have developed
various different technologies. Kyocera has worked with other Japanese companies, for
example, Osaka Gas Co., Ltd., to produce a ‘flat tubular’ SOFC that is comprised of a
series of parallel tubes within one ceramic tablet. This robust configuration, shown in
Figure 9.14, is used as the core fuel‐cell technology for a number of Japanese micro
combined heat and power (CHP) units. The strategy of MHPS, by contrast, is to
255
256
Fuel Cell Systems Explained
Figure 9.13 Acumentrics tubular SOFC bundle as used in the RP1500, 1.5‐kW system.
(Source: Reproduced with permission of Acumentrics.)
Figure 9.14 (a) Kyocera–Osaka gas power‐generation unit (left) and heating unit that uses SOFC
exhaust heat (right) and (b) Kyocera flat tubular cells. (Source: Reproduced with permission of Kyocera.)
implement SOFC stacks in combined‐cycle or hybrid systems, as described in Section 9.6.
Both the Kyocera and MHPS stacks are designed to be fuelled with natural gas.
Mitsubishi Heavy Industries, Ltd. (Nagasaki) and Tokyo Gas Co., Ltd. have also
collaborated on tubular SOFC technology. In 2003, a ‘flat tubular’ SOFC built by NGK
Insulators demonstrated a power density of 0.6 W cm−2 at 650°C and 1.6 W cm−2 at
750°C — a world record at that time. TOTO started a programme on tubular SOFCs in
1989 and by 2001 had produced 10‐kW stacks. The TOTO fuel cells were similar in
appearance to those of Westinghouse tubular SOFCs, but the cell components were
deposited using a wet chemical process rather than by EVD.
9.5.2
Planar SOFCs
The organizations currently involved in the research, development and commercialization of planar SOFCs are too numerous to list here. Much of the early work
Solid Oxide Fuel Cells
was undertaken in the United States to be followed in the 1980s by that conducted
in Europe and Japan. Some of the early players in the United States, such as Allied
Signal Aerospace Co. and SofCo EFS,9 are no longer in the fuel‐cell business,
whereas the others, such as Versa Power Systems, Inc., Ceramatec Inc. and Delphi
Automotive LLP, which were supported under the SECA programme, are still active.
In terms of commercialization status, the major planar SOFC provider in the
United States is now Bloom Energy. This company was founded in 2001 with the
name Ion America and is based in California (USA). It changed its name in 2006
after attracting US$400 million in investments. In the same year, Bloom shipped its
first 5‐kW demonstration unit to the University of Tennessee, Chattanooga. After
2 years of field trials in Tennessee, California and Alaska, the first pre‐commercial
prototype product was shipped to Google in July 2008. The company has continued
increasing the size of their systems during the last few years through the production
of the following models of Bloom’s Energy Server®: ES‐5000, ES‐5400 and ES‐5700
that generate 100, 105 and 210 kW, respectively. Each power generator is built up
with 1‐kW stacks that are composed of 40 × 25‐W cells, fuelled with natural gas.
Stacks are combined to provide a given power output, and the system is marketed
as a ‘Bloom Box’. Many systems have been installed for eminent clients throughout
the United States. In January 2011, Bloom started to offer a bold innovative service
called ‘Bloom Electrons’ that would allow customers to purchase the electricity
provided by the Bloom Box at a set price for 10 years without incurring any
other costs.
In Japan, developers of planar SOFCs include Fuji Electric, Co., Ltd.; Tokyo Gas, Co.,
Ltd.; Mitsubishi Heavy Industries, Ltd.; Mitsui Engineering and Shipbuilding, Co., Ltd.;
Murata Manufacturing, Co., Ltd.; Sanyo Electric, Co., Ltd.; Tokyo Gas Co., Ltd.; and
Tonen. In Europe, innovators of planar SOFCs with good track records have been a
consortium of Haldor Topsoe A/S and Riso National Laboratory (Denmark),
Forschungszentrum Jülich in collaboration with Sunfire GmbH (Germany), Wärtsilä
Corporation (Finland), Hexis Ltd. (Switzerland) and Ceres Power (United Kingdom).
There are of course many differences between the various technologies under
development. Ceres Power, for example, is focused on an IT‐SOFC known as the ‘Steel
Cell’. Whereas many planar cells use anode‐supported structures, the components in
the Steel Cell are deposited directly onto a porous stainless‐steel substrate. Intended to
be fuelled by natural gas, the cell operates in the range 500–600°C, is able to be started
quickly and will endure thermal cycling better than ceramic‐supported SOFCs. Ceres
Power is currently supplying stacks to original equipment manufacturers (OEMs)
partners in the United Kingdom, Korea and Japan.
The Rolls‐Royce ‘Integrated Planar’ SOFC is an especially distinctive technology.
Work on the concept started in Derby, UK, in the mid‐1980s. From the outset, the aim
was to target low‐cost manufacturing that could enable affordable MW‐class systems to
be built. The Rolls‐Royce cells are screen printed onto a porous extruded ceramic
substrate that gives a narrow cell pitch to reduce the ohmic losses; a schematic diagram
of the arrangement is given in Figure 9.15. Steady progress was made in the United
Kingdom during the 1990s, and by 2001 a 300‐mm, 40‐cell proof‐of‐concept module
generated 27.3 W and achieved a power density of 155 mW cm−2 at an average cell
9 SOFCo EFS Holdings LLC was acquired by Rolls‐Royce Fuel Cells from McDermott International in 2007.
257
258
Fuel Cell Systems Explained
Anode
current-collector
Cathode
Air
Microporous
barrier layer
YSZ
Anode
Interconnect
Fuel
Porous ceramic substrate
Figure 9.15 Rolls‐Royce Integrated Planar SOFC concept.
voltage of 0.62 V with 43% utilization of simulated reformed fuel gas. In June 2012, LG
Electronics Inc., a South Korean company, acquired the Rolls‐Royce fuel‐cell business,
which is now known as LG Fuel Cell Systems Inc.
A final mention should be given to SOLIDpower S.p.A., an Italian company that was
formed in 2006 following the acquisition of HTCeramix SA, a private company in
Yverdon‐les‐Bains, Switzerland, that had commenced work on SOFCs four years earlier.
The latter was a spin‐off of the Swiss Federal Institute of Technology in Lausanne
(EPFL). In 2015, SOLIDpower S.p.A. also acquired the European assets and employees
of Ceramic Fuel Cells Ltd. (CFCL), an Australian SOFC developer that, through lack of
funding, had ceased trading earlier in the year. The company, which was a spin‐off
in 1992 from the Commonwealth Scientific and Industrial Research Organisation
(CSIRO), had produced for the domestic market a packaged CHP system (2.5 kWe,
2.0 kWth) — the ‘BlueGen’ — fuelled by natural gas.
9.6
Combined‐Cycle and Other Systems
It has been mentioned in previous chapters that a high‐temperature fuel cell can be
combined with a steam turbine in a bottoming cycle. The ability to use both gas turbines
and steam turbines in a combined cycle with an SOFC has been known in concept for
many years. It is only recently, however, that pressurized operation of SOFC stacks has
been demonstrated for prolonged periods, thereby making the SOFC–gas turbine
(SOFC–GT) system feasible practically. Pioneered by Siemens Westinghouse Power
Corporation in their SureCell™ concept, the ideas of combined SOFC–GT are now
being explored by other developers such as Mitsubishi Heavy Industries; the essential
process features are illustrated diagrammatically in Figure 9.16.
In this review of the status of SOFCs, it is worthwhile remarking that there are many
opportunities for novel system design, as well as scope for considerable creativity by the
systems engineer. Many examples have been reported in the literature. For instance, by
connecting stacks in series, a multistage SOFC — the ‘UltraFuelCell’ — was developed
under a programme supported by the US Department of Energy. A novel hybrid
system concept has also been described in which SOFC and PEMFC technologies are
Solid Oxide Fuel Cells
After
burner
Natural
gas
C
a
t
h
o
d
e
A
n
o
d
e
Air
Gas
turbine
Alternator
(200 kW)
Fuel cell
(800 kW)
Exhaust
Preheaters
Desulfuriser
Figure 9.16 System concept for an SOFC–GT combined cycle.
combined.10 The advantages of each type of fuel cell are enhanced by operating in
synergy; the system is shown in Figure 9.17. The IT‐SOFC is run under conditions that
give low fuel utilization and thereby enables a high‐power output for a relatively small
stack size. Unspent reformed fuel appears in the anode exhaust where it undergoes a
shift reaction, followed by a process stage when the final traces of carbon monoxide are
removed. At this stage, the gas comprises mainly hydrogen and carbon dioxide, with
some steam. This gas, once it is cooled, is suitable as a fuel in the PEMFC stack. The gas
compositions for the system are listed in Table 9.1. The use of two stacks of different
types for power generation results in increased overall electrical efficiency. The system
is particularly attractive in terms of economics. Preliminary calculations reveal that the
system is more competitive than an SOFC‐only system because of the anticipated
relatively low cost of the PEMFC stack. On the other hand, the system would offer a
much higher efficiency than could be achieved by a PEMFC‐only system when operating
on natural gas. In the following chapter, it is shown that the fuel‐processing technology for
running a PEMFC on natural gas is complex, bulky and expensive. How much better, then,
to use an SOFC as the fuel processor!
10 Dicks, AL, Fellows, RG, Mescal, CM, Seymour, C 2000, A study of SOFC–PEM hybrid systems, Journal
of Power Sources, vol. 86(1–2), pp. 501–506.
259
260
Fuel Cell Systems Explained
Exhaust
air
Fuel in
(e.g. methane)
Solid oxide
fuel cell
Shift
reactors
Selective
oxidation
Air
in
PEM
fuel cell
Burner
Exhaust
air
Heat addition
to system
Heat removal
from system
Figure 9.17 SOFC–PEMFC hybrid system (see also Table 9.1).
Table 9.1 Summary of output powers for the hybrid system
shown in Figure 9.17.
SOFC stack power (kW)
369.3
PEM stack power (kW)
146.7
Turbine power (kW)
Compressor power (kW)
100.3
–100.8
Net power output (kW)
515.5
Electrical output (kW)
489.7
Overall efficiency, % (LHV)
61
Further Reading
Atkinson, A, Barnett, S, Gorte, RJ, Irvine, JTS, McEvoy, AJ, Mogensen, M, Singhal, SC and
Vohs, J, 2004, Advanced anodes for high‐temperature fuel cells, Nature Materials, vol. 3,
pp. 17–27.
Cowin, P, Petit, C, Lan, R, Irvine JTS and Tao, S, 2011, Recent progress in the development
of anode materials for solid oxide fuel cells, Advanced Engineering Materials, vol. 1,
pp. 314–312.
Solid Oxide Fuel Cells
Fergus, JW, 2005, Metallic interconnects for solid oxide fuel cells, Materials Science and
Engineering: A, vol. 397, pp. 271–283.
Irvine, JTS, Neagu, D, Verbaeke, MC, Chatzichristodoulou, C, Graves, C and Mogensen,
MB, 2016, Evolution of the electrochemical interface in high‐temperature fuel cells and
electrolysers, Nature Energy, vol. 1, 15014, available online at http://palgrave.nature.
com/articles/nenergy201514 (accessed 24 September 2017).
Oishi, N, Rudkin, RA, Steele, BCH and Brandon, NP, 2002, Thick Film Stainless Steel
Supported IT‐SOFCs for Operation at 500‐600°C, Scientific Advances in Fuel Cells,
Elsevier Science Ltd, Amsterdam.
Singhal, SC, 2007, Solid oxide fuel cells, The Electrochemical Society, Interface, Winter
2007, pp. 41–44.
Singhal, SC and Kendall, K (eds.), 2003, High Temperature Solid Oxide Fuel
Cells – Fundamentals, Design and Application, Elsevier, B.V, Amsterdam.
Steele, BCH and Heinzel A, 2001, Materials for fuel cell technologies, Nature, vol. 414,
pp. 345–352.
Wei, T, Singh, P, Gong, Y, Goodenough, J, Huang, Y and Huang, K, 2014, Sr3‐3x Na3x
Si3O9‐1.5x (x=0.45) as a superior solid oxide ion electrolyte for intermediate temperature
solid oxide fuel cells. Energy and Environmental Science, vol. 7, pp. 1680–1684.
261
263
10
Fuels for Fuel Cells
10.1
Introduction
This chapter considers fuels that can be used for the principal types of fuel cell. Hydrogen
has been promoted worldwide as a panacea for energy problems in that it may eventually
replace, or at least greatly reduce, the reliance on fossil fuels while being itself a
clean‐burning fuel that releases no greenhouse gases into the atmosphere. Although
the most abundant element in the universe — the stuff from which stars are made —
hydrogen does not occur freely on earth, but is predominantly found in combination
with oxygen as water and with carbon as fossil fuels. Chemical, thermal or electrical
energy has to be expended to extract hydrogen from these sources. Hydrogen is
therefore not a new form of primary energy, but a vector (or carrier) for storing and
transporting energy from any one of a myriad of sources to where it may be utilized.
Basic chemical and physical data on hydrogen and some of the other fuels that may be
suitable for fuel cells are given in Table 10.1.
Hydrogen has the advantage over all other fuels in that it is easily oxidized at the
fuel cell anode and that the only chemical product is water. For this reason, hydrogen
has become preferred for fuel‐cell vehicles (FCVs). Cars and other vehicles that
employ hydrogen‐fuelled proton‐exchange membrane fuel cells (PEMFCs) are
‘zero emission’, because the only exhaust from the vehicle tailpipe is water or
water vapour.1
Various primary fossil fuels that can be used to generate hydrogen are discussed in
some detail in Section 10.2. These hydrocarbon fuels include natural gas, petroleum
products (such as gasoline and diesel), coal gas and coal. Biofuels are also a possible
source of hydrogen, and the options for using such fuels are reviewed separately in
Section 10.3. The processes and chemical‐conversion technologies for obtaining hydrogen
from hydrocarbons, whether these are from fossil fuels or biofuels, are well established
industrially and are covered in Sections 10.4–10.6. For stationary fuel‐cell power plants,
there are strong arguments for producing hydrogen as close to the fuel‐cell stack as
possible. This is because, apart from safety considerations, the heat generated in the fuel‐
cell stack may then be conveniently used for some of the fuel processing. Integration of
1 In a combustion engine or gas turbine, not only is hydrogen oxidized to water (or steam), but also
nitrogen in the air is oxidized to yield nitrogen oxides (NOx).
Fuel Cell Systems Explained, Third Edition. Andrew L. Dicks and David A. J. Rand.
© 2018 John Wiley & Sons Ltd. Published 2018 by John Wiley & Sons Ltd.
264
Fuel Cell Systems Explained
Table 10.1 Some properties of hydrogen and other candidate fuels for fuel‐cell systems.
Hydrogen
(H2)
Molecular weight (g)
2.016
Methane
(CH4)
16.04
Ammonia
(NH3)
17.03
Methanol
(CH3OH)
32.04
Ethanol
(C2H5OH)
46.07
Gasoline
(C8H18)a
114.2
Freezing point (°C)
−259.2
−182.5
−77.7
−98.8
−114.1
–56.8
Boiling point (°C)
−252.77
−161.5
−33.4
64.7
78.3
125.7
Net enthalpy of
combustion at 25°C
(kJ mol−1)
241.8
802.5
316.3
638.5
1275.9
5512.0
Heat of vapourization
(kJ kg−1)
445.6
510
1371
1129
839.3
368.1
Liquid density (kg L−1)
77
425
674
786
789
702
Specific heat (STP)
(J mol−1 K−1)
28.8
112.4
188.9
Flammability limits
in air (%)
Auto‐ignition
temperature in air (°C)
a)
34.1
36.4
76.6
4–77
4–16
15–28
6–36
4–19
1–6
571
632
651
464
423
220
Gasoline is a blend of hydrocarbons and varies with producer, application and season. N‐octane is
reasonably representative of properties except for vapour pressure, which is intentionally raised by
introducing light petroleum fractions.
the fuel‐cell stack and fuel processor is therefore an important aspect of system design, as
outlined in Section 10.5.
In the early days of FCV development, some of the concept cars and buses ran on
methanol (CH3OH) — an on‐board reformer converted the methanol into a hydrogen‐
rich gas for the fuel cell. Even on‐board reforming of gasoline was investigated with
support from the US Department of Energy (DOE). In recent years, however, the
practice has largely fallen out of favour because such generation of hydrogen can be a
complex and not particularly efficient process. Moreover, dispensing with on‐board
fuel processing greatly simplifies the design of the vehicle drivetrain and ensures that
water is the only emission. The implications for moving from on‐board to off‐board
reforming of fuel are discussed in Section 10.7.
Hydrogen may, of course, be generated by the electrolysis of water — the reverse of a
fuel cell. Since the purpose of a fuel cell is to produce electricity, this may at first seem
perverse. In many cases, however, electrolysis is a very convenient means of providing
hydrogen for small mobile fuel cells and even vehicles. Attention is currently being
given to the large‐scale generation of hydrogen by electrolysis as a means of storing
renewable energy. To this end, several experimental power‐to‐gas (P2G) schemes have
been built in which solar or wind power is used to produce hydrogen by electrolysis, as
reported in Section 10.8.
Thermochemical methods of hydrogen generation are evaluated in Section 10.9. The
chapter concludes (Section 10.10) with an appraisal of hydrogen generation by biological
systems that employ processes that involve enzymes, bacteria and sunlight.
Fuels for Fuel Cells
Fossil fuels
(e.g., gas, oil, coal;
Section 10.2)
Large chemical
plants reforming
fuels to hydrogen
(Section 10.4)
Hydrogen stored as
gas at pressure or
cryogenic liquid
(Section 11.3 and 11.4)
Renewable fuels
(e.g., biofuels, waste;
Section 10.3)
Biological hydrogen
generation systems
(Section 10.10)
Hydrogen stored as synthetic
fuel—methanol, ammonia,
sodium borohydride
(Sections 11.5 and 11.6)
Central or dispersed
power generation
(Chapter 12)
Thermochemical
hydrogen
production
(Section 10.9)
Electrolysis
(Section
10.8)
Small-scale hydrogen
storage (e.g., high pressure
cylinder or metal hydride
cannister;
Section 11.7)
Small-scale fuel
reformer producing
hydrogen
(Section 10.6)
Stationary fuel-cell
system with MCFC
or SOFC using fuel
directly
(Chapters 8 and 9)
Electricity generated by
renewable energy, solar,
wind, wave, hydro, etc,
Nuclear
Stationary fuel-cell
system using
hydrogen
Stationary or
portable fuelcell system
using
synthetic fuel
Hydrogen fuel-cell
system for vehicles
(Chapter 4)
Figure 10.1 The many different pathways by which hydrogen can be supplied to fuel cells.
From the aforementioned overview, it will be evident that there are many possible ways
of producing hydrogen for a fuel cell. An illustration of how the different fuel‐processing
pathways are related is presented in Figure 10.1, together with directions as to the
appropriate information that can be found in this chapter and others.
265
266
Fuel Cell Systems Explained
10.2
Fossil Fuels
10.2.1
Petroleum
Petroleum is a mixture of gaseous, liquid and solid hydrocarbon‐based chemical
compounds that occur in sedimentary rock deposits throughout the world. Crude
liquid petroleum (crude oil) has little value, but when refined, it provides high‐value
liquid feeds, solvents, lubricants and other products. Petroleum‐derived fuels account
for up to one‐half of the world’s total energy supply and include gasoline (petrol),2 diesel
fuel, aviation fuel and kerosene. Simple distillation is able to separate various components
of crude petroleum into generic fractions with different boiling ranges, as shown in
Figure 10.2. The amounts of the fractions that are obtained from any particular variety
of crude oil depend on the origin of supply.
Each fraction of petroleum contains a different proportion of chemical compounds,
these being normal and branched paraffins or alkanes, monocyclic and polycyclic
C1 to C4
20°C
C5 to C3
Fractions
decreasing in
density and
boiling point
70°C
C5 to C10
120°C
C10 to C16
Fractionating
column
170°C
Liquefied petroleum gas
Chemicals
Petrol for vehicles
Jet fuel
Paraffin for lighting and heating
Diesel fuels
C14 to C20
270°C
Crude oil
C20 to C50
Fractions
increasing in
density and
boiling point
Lubricating oils,
waxes, polishes
C20 to C70
Fuels for ships, factories
and central heating
600°C
>C70 residue
Bitumen for
roads and roofing
Figure 10.2 Distillation products of crude petroleum.
2 Colloquially the word ‘petrol’ is used to describe the fuel for cars in the United Kingdom, Australia and
other countries, whereas the term ‘gasoline’ is more commonly used for the same fuel in the United States.
Fuels for Fuel Cells
paraffins (naphthenes) and mononuclear and polynuclear aromatic hydrocarbons. The
naphtha fraction (C7–C11) principally comprises normal alkanes and some monocyclic
alkanes (e.g., cyclopentane and cyclohexane). Low boiling fractions generally have more
of these low molecular weight alkanes than the high boiling fractions. Similarly the
content of polycyclic alkanes and aromatic hydrocarbons increases on moving from low
boiling to high boiling fractions. The various fractions can be used without further
refining and are referred to as ‘straight‐run’ fractions. In the early days of motoring, for
example, gasoline was a straight‐run distillate. As demand for fuel increased, more
gasoline (which typically comprises alkanes with between 4 and 12 carbon atoms in
each molecule) was produced by chemically cracking the fractions that contained
hydrocarbons of high molecular weight and blending the product with the straight‐run
distillate. Nowadays, fuels for vehicles are blended from several petroleum refinery
process streams that are derived by the following methods: direct distillation of crude
oil, catalytic and thermal cracking, hydrocracking, catalytic reforming, alkylation and
polymerization. In addition, gasoline can be doped with various compounds to improve
engine lubrication, reduce corrosion and lessen the risk of ‘knock’ or pre‐ignition in
internal combustion engines.3 Other engine performance enhancers that may be added
include antioxidants, metal deactivators, lead scavengers, anti‐rust agents, anti‐icing
agents, detergents and dyes. At the end of the refining process, finished gasoline
typically contains more than 150 separate compounds although as many as 1000
compounds have been identified in some blends. Clearly gasoline and other petroleum
products can be chemically complex mixtures, but further discussion of their production
and composition is outside the scope of this book.
The importance of discussing the composition of vehicle fuels is that the choice of the
process required for generating hydrogen is determined by the chemical composition
and the physical and combustion characteristics of the fuel. In the case of catalytic
conversion, the presence of various trace compounds in the fuel may be of serious
concern as they may act as poisons for the conversion catalysts and indeed also for
anode catalyst in the fuel‐cell stack. In this context, the problematic trace compounds
in fossil fuels are organic compounds that contain silica and sulfur, and organometallic
compounds, such as various porphyrins. Sulfur is a persistent poison for catalysts and
its removal from raw fuels is discussed specifically in Section 10.4.2.
The gasoline fraction of petroleum, together with the heavier diesel, is widely distributed
as a fuel for vehicles that range from passenger cars to heavy‐duty trucks, buses and rail
locomotives. Because the infrastructure for delivering such fuels is mature, there are
valid reasons for using these fuels for FCVs. Components that are currently added to
gasoline or diesel for anti‐knock or lubrication purposes would not be required for
such vehicles so that future refinery operations could be simplified. Techno‐economic
studies and ‘well‐to‐wheel’ analyses indicate, however, that the benefits of using
gasoline in FCVs, compared with internal combustion engines, are less than those when
operating with hydrogen or even methanol (cf. Section 12.4, Chapter 12).
3 Lead tetraethyl, (CH3CH2)4Pb, was added to gasoline or petrol in the 1920s to improve the octane rating
to avoid pre‐ignition. Such leaded fuel was gradually phased out from the 1970s because of concern over its
toxicity. Modern refining operations can produce an unleaded gasoline with a sufficiently high octane rating
(98 octane) suitable for all high‐compression engines.
267
268
Fuel Cell Systems Explained
Progressively more demanding emissions standards for vehicles have led to lower
sulfur contents in distributed gasoline and diesel fuels in both Europe and the United
States. This trend is expected to continue and it may lead to more widespread use of
synthetic fuels, or biofuels. Synthetic diesel fuel made, for example, by the transesterification of vegetable oils or animal fats is intrinsically low in sulfur. Other biological
sources are being investigated for future alternative fuels, notably algae and some lipids
from waste waters. Biofuels can be used not only for land‐based vehicles, such as cars,
buses, trucks and trains, but also for aircraft. Vehicles running on biofuels offer a
demonstrable method of reducing carbon emissions.
Other hydrocarbons of low molecular weight, such as propane (C3H8) and butane
(C4H10), are often found associated in crude oil deposits. When this material is distilled,
these hydrocarbons emerge as a gaseous product, which is marketed as liquefied
petroleum gas (LPG). Available therefore as a by‐product from oil refineries, LPG is a
gas at atmospheric pressure that will liquefy at a moderately elevated pressure and thus
can be transported easily. The fuel is used extensively for applications as diverse as
camping gas stoves and vehicle propulsion. In addition, LPG is attractive for fuel‐cell
systems that may be used as remote stationary power sources where there is no pipeline
source of gas and possibly also for some FCV applications.
10.2.2
Petroleum from Tar Sands, Oil Shales and Gas Hydrates
Petroleum substances can be found in the Earth’s crust as deposits of solid or near‐solid
material in sandstone at depths that are usually less than 2000 m. Some may also be
found as an outcrop on the surface. Huge amounts of such ‘tar sands’ are present in
various parts of Canada and the United States, but the high bitumen content (of very
high molecular weight) makes recovery less attractive than from more conventional
petroleum deposits. Similarly, ‘oil shales’ comprise a significant and largely untapped
source of petroleum materials. Oil shales are compact laminated rocks of sedimentary
origin in which petroleum is locked. The oil can be obtained from the rock by distillation.
Global deposits of shale oil are estimated to be around 5 × 1012 barrels, that is, significantly
more than crude oil resources, which stand at only 1.5 × 1012 barrels.4 Compared with
crude oil, however, only a small fraction of oil from shales is easily recoverable by
conventional technologies; much of the oil is of high molecular weight and bituminous
in nature. While oil shales may become an important energy resource into the future, a
discussion of the processing of such materials is outside the scope of this book.
In natural petroleum reservoirs where the prevailing pressures are high and
temperatures low (e.g., under permafrost), methane (CH4) and other normally gaseous
hydrocarbons form ice‐like hydrogen‐bonded complexes with water that are known as
‘gas hydrates’. Methane hydrates have been considered as a method of transporting
natural gas from remote fields.
10.2.3
Coal and Coal Gases
Coal is the most abundant of all fossil fuels. Chemically, the resource is the most complex
given that it is formed from the compaction and induration of various plant remains
4 World Energy Resources, 2013, World Energy Council. ISBN: 978 0 946121 29 8.
Fuels for Fuel Cells
that were laid down in geological time, especially during the Carboniferous Period
between 345 and 280 million years ago. The first product of decay and consolidation is
peat, which has a relatively low carbon content and a high moisture content. Under
forces of heat and pressure, peat gradually converts first into bituminous coal and,
ultimately, to hard coal (anthracite). The diversity of the original plants, the variations
in the depositional environment and the age of the coal since it was laid down (the ‘rank’
of the coal) have resulted in an exhaustive literature that classifies coal by its appearance
(macroscopic and microscopic), its chemical composition, the occluded mineral matter,
and its physical properties.
It is worth pointing out that apart from combustion, further processing of coal to
produce liquids, gases and coke is highly dependent on the properties of the raw coal
material. For example, coals with 20–30 wt.% volatile organic matter are primarily suitable
for producing coke. By contrast, bituminous coals are best suited for carbonization, i.e.,
heating to temperatures of 750–1500°C to form ‘coal gas’, also known as ‘town gas’. Most
types of coal can be gasified, although the rank and other characteristics of the coal will
influence the product mix from the various designs of gasifier.
Coal carbonization was the original method for producing coal gas in the 19th century
in both the United Kingdom and North America. Simple carbonization, i.e., the
destructive distillation of coal in the absence of air, yielded gas (a mixture mainly of
hydrogen and carbon oxides), organic liquids (tars and phenolics) and a residual coke.
Partial pyrolysis of coal was also carried out in coke ovens where the main objective was
to produce coke for the steel industry. Such ovens generated combustible gas that could
also be used in industry. Another type of coal gas, known as ‘producer gas’, was obtained
by blowing air and steam over hot coke at high temperatures. Most of these gas‐making
processes are now obsolete. In the 1950s, methods of manufacturing gas from oil were
developed in the United Kingdom and the United States. The production of coal gas by
the so‐called catalytic rich gas (CRG) process had been proven at the pilot scale by
engineers in the United Kingdom, but with the advent of natural gas from the North
Sea, further development was curtailed. Nevertheless, the CRG process is still operated
in various locations around the world where manufactured gas is required.
Nowadays, coal carbonization has been superseded for large‐scale gas production by
various coal gasification processes. Gasification differs from carbonization in that the
heated coal is reacted with steam and oxygen (or air) at high temperatures. The products
of primary coal gasification are mainly gases, together with smaller amounts of liquids
and solids. The relative proportion of products depends on the type of coal, the
temperature and pressure of reaction and the relative amounts of steam or oxygen that
are injected into the gasifier. Coke is not formed and the only waste material is an inert
ash or slag. Further chemical processing of the raw gas can be carried out, for example,
to increase the CH4 content or to alter the hydrogen‐to‐carbon monoxide (CO) ratio to
suit the final application for the coal gas.
The numerous gasification systems available today can be broadly classified into three
basic types of gasifier: (i) moving bed, (ii) entrained bed and (iii) fluidized bed. The moving‐
bed gasifiers produce a gas at low temperature (450–650°C) that contains CH4 and
ethane (C2H6), which arise from devolatilization of the coal, together with a hydrocarbon
liquid stream that contains naphtha, tars, oils and phenolic liquids. Entrained‐bed
gasifiers produce gas at higher temperatures (>1200°C) with virtually no hydrocarbons
in the resultant gas stream and much lower amounts of liquid hydrocarbons. In fact, the
269
270
Fuel Cell Systems Explained
Table 10.2 Typical coal gas compositions (mole per cent basis).
BG–Lurgi
gasifier
(non‐
slagging)
BG–Lurgi
slagging
gasifier
Moving‐bed
O2‐blown
(BG–Lurgi)
Fluidized‐
bed
(Winkler)
Coal
Pittsburgh
No.8 seam
Pittsburgh
No.8 seam
Illinois
No.6 trace
Illinois
Texas lignite Illinois No.6 Illinois No.6 No.6
Ar
Trace
Trace
Trace
CH4
9.43
6.55
3.26
Entrained‐
bed O2‐
blown
(Texaco)
0.90
1.10
5.93
0.12
Entrained‐
bed air‐
blown
Trace
1.08
Entrained‐
bed O2‐
blown
(Shell)
1.15
0.00
C2H6
0.78
0.09
0.10
0.00
0.00
0.00
0.00
C2H4
0.33
0.18
0.20
0.00
0.00
0.00
0.00
H2
32.30
35.49
20.75
36.47
37.18
9.68
27.99
CO
19.98
50.50
5.73
42.65
48.59
17.20
66.14
CO2
34.52
3.55
11.66
19.97
13.25
6.45
1.57
2.66
3.64
0.20
0.77
0.86
66.67
4.30
N2
NH4
0.00
0.00
0.40
0.13
0.12
0.00
0.00
H2O
0.00
0.00
61.07
21.65
20.25
5.38
2.10
H2S
Total
0.00
0.00
0.49
0.26
1.23
0.00
1.36
100.00
100.00
100.00
100.00
100.00
100.00
100.00
entrained‐bed gas is composed almost entirely of hydrogen, carbon monoxide and
carbon dioxide (CO2). In terms of composition and temperature (925–1050°C), the gas
from the fluidized‐bed gasifier falls somewhere between those obtained from the
two other designs of reactor. For all three gasifiers, the heat required for the reaction of
coal and steam is effectively provided by the partial oxidation (POX) of the coal. The
temperature, and therefore the composition, of the gas is dependent on the proportions
of oxidant and steam, as well as the design of the reactor. The chemical compositions of
typical coal gases from some of the leading types of gasifier are given in Table 10.2. The
gases invariably contain contaminants that must be removed before use in fuel cells.
Methods of gas clean‐up are described in the next section.
10.2.4
Natural Gas and Coal‐Bed Methane (Coal‐Seam Gas)
Natural gas is combustible and occurs in porous rocks such as sandstone in the Earth’s
crust. It is found with, or close to, crude oil reserves but may also be present in separate
reservoirs. Most commonly, it is trapped between liquid petroleum and an impervious
rock layer (‘cap rock’) in a petroleum reservoir. If the pressure is sufficiently high, the
gas will be intimately mixed with, or dissolved in, the crude oil.
Natural gas is comprised of a mixture of hydrocarbons of low boiling point. Usually,
methane is present in the greatest concentration, with smaller amounts of ethane,
propane and higher‐order alkanes. In addition to hydrocarbons, natural gas contains
Fuels for Fuel Cells
Table 10.3 Typical compositions of natural gases from different geographic regions.
Component
North Sea
Qatar
the Netherlands
Pakistan
Ekofisk,
Norway
Indonesia
CH4
94.86
76.6
81.4
93.48
85.5
84.88
C2H6
3.90
12.59
2.9
0.24
8.36
7.54
2.38
0.4
0.24
2.85
1.60
0.04
0.86
0.03
0.1
0.06
C3H8
i‐C4H10
0.15
n‐C4H10
C5+
N2
S
0.11
0.21
0.02
0.79
0.24
14.2
4 ppm
1.02
1 ppm
0.12
0.41
0.22
1.82
4.02
0.43
4.0
30 ppm
2 ppm
N/A
Values are vol.%, unless otherwise stated.
various quantities of nitrogen and carbon dioxide (CO2), together with traces of other
gases such as helium (often present in commercially recoverable quantities). Sulfur is
also present to a greater or lesser extent, mostly in the form of hydrogen sulfide (H2S).
Natural gas is often described as being dry or lean (containing mostly CH4), wet
(containing considerable concentrations of higher molecular weight hydrocarbons),
sour gas (with significant levels of H2S), sweet gas (low in H2S) and casinghead gas
(obtained from an oil well by extraction at the surface). Some typical compositions of
natural gases from different regions of the world are listed in Table 10.3.
Coal‐bed methane (CBM), sometimes known as coal‐seam gas (CSG), is found
absorbed in the solid matrix of underground deposits of coal. It is only in recent years
that CBM has been recognized as a huge energy resource. In Queensland, Australia, for
example, investment of over US$20 billion has been spent on the infrastructure to
extract the gas from vast areas of underground coals and then combine and transfer the
outputs from different wells to a liquefied natural gas (LNG) terminal. Located in
Gladstone, the LNG terminal has been built to supply a growing demand in Asia for the
fuel. Unlike natural gas, CBM is usually very low in sulfur and other hydrocarbons such
as ethane and propane. Sometimes, there are significant concentrations of CO2 and N2
associated with CBM that may have to be removed before the gas is transported. The
chief problem, however, lies with the water that is brought to the surface when the gas
is released from the coal. This water has to be separated and dispatched via irrigation,
evaporation or reinjection into underground reservoirs.
Some processing of natural gas or CBM may therefore be necessary close to the
point of extraction before it enters a transmission system. Examples with natural gas
are the bulk removal of sulfur (sweetening) and the removal of high molecular weight
hydrocarbons, nitrogen, acid gases, liquid water and liquid hydrocarbons. With
CBM, the only processing usually required is the removal of nitrogen or acid gases
and water.
There are wide variations in the composition of natural gas fed to transmission
systems around the world, as indicated in Table 10.3. Even within geographic regions
271
272
Fuel Cell Systems Explained
and also according to the season, there may be significant variations in composition
according to the field from which the gas is obtained. In colder months, the heating
value of the gas may drop due to a fall in concentration of higher hydrocarbons. At such
times it is common practice to enrich the gas by blending in mixtures of ethane, propane
and butane to ensure that the distributed gas has consistent burning characteristics and
heating value throughout the year. This raises a fundamental problem for developers of
fuel cells. That is, the design of a fuel process is influenced by the composition of the gas
rather than by its combustion properties. The situation is exacerbated by gas companies
that may enrich the gas with mixtures of propane–air or butane–air in order to
boost the heating value during times of seasonal peak demand. Certainly, reforming
catalysts — such as the Johnson Matthey CRG catalysts (v.s.) widely used in the steam
reforming of natural gas (see Section 10.4.3) — will only tolerate a very small percentage
of oxygen in the feed gas.
Natural gas has no distinctive odour (except for very sour gases), and for safety
reasons, therefore, pipeline companies and utilities usually odourize the gas5 either as it
enters the transmission system or within local distribution zones. Various odourants
may be used and the most common are thiophenes and mercaptans. Tetrahydrothiophene
(THT) is widely used throughout Europe and in the United States (where it is known as
Pennwall’s odourant), whereas in the United Kingdom and Australia, for example, a
cocktail of compounds is used (a combination of ethyl mercaptan, tertiary butyl
mercaptan and diethyl sulfide).
Coal‐bed methane is quite different from natural gas in that the composition (typically
greater than 92% CH4 on a dry basis) is remarkably consistent throughout the lifetime
of the well. The gas is also naturally low in sulfur, and therefore once any associated
water or acid gases are removed, only the odourant that may be injected is likely to be a
concern of fuel‐cell engineers.
10.3
Biofuels
Biomatter (or biomass) is a catch‐all term for the natural organic material associated
with living organisms that include terrestrial and marine vegetable matter — everything
from algae to trees — together with animal tissue and manure. On a global basis, it is
estimated that over 150 Gt of vegetable biomatter is generated annually. The production
of biomass material is often expressed in tonnes per hectare, (t ha−1); 1 ha = 104 m2. This
yield ranges from about 13 t ha−1 for water hyacinth to 120 t ha−1 for Napier grass. In
view of its high energy content, biomass represents an important source of a renewable
fuel that can be obtained via the following routes:
●
●
●
●
●
Direct combustion.
Conversion to biogas via pyrolysis, hydrogasification or anaerobic digestion.
Conversion to ethanol (C2H5OH) via fermentation.
Conversion to syngas thermochemically.
Conversion to liquid hydrocarbons by hydrogenation.
5 LPG is also odourized, usually with ethyl mercaptan.
Fuels for Fuel Cells
Box 10.1 Syngas
Syngas (also known as ‘synthesis gas’) is the product gas from a steam reformer and
consists primarily of hydrogen, carbon monoxide and often some carbon dioxide. In fact,
the term can be used for any gas mixture that contains hydrogen and carbon monoxide,
e.g., the gas produced by gasification of coal. The name refers to its use in the synthesis of
substitute (or synthetic) natural gas (SNG) and ammonia or in the production of synthetic
liquid hydrocarbon fuels by the Fischer–Tropsch process, e.g., synthetic diesel and
synthetic gasoline. The latter is now more commonly known as the ‘gas‐to‐liquid (GTL)’
process and is carried out over a nickel‐, cobalt‐ or thorium‐based catalyst at 200°C, i.e.,
nCO
2n 1 H2
C nH2 n
2
nH 2 O
Another source of biofuel is municipal waste, i.e., sewage sludge and municipal solid
waste (MSW). The latter is a general term applied to solid household or commercial
garbage. Methane has been produced by the biological digestion of sludge in waste‐water
treatment plants for many years. In many cases, the methane is used to fuel generators
that provide supplementary power to assist the running of the treatment plant. Fuel‐cell
systems have been employed for such power generation in several demonstrations
throughout the Europe Union (EU) and the United States.
Gaseous fuels that emanate from landfill sites and other forms of refuse digestion can
also constitute a useful source of energy that is well suited to fuel‐cell systems. According
to the European Commission, in the EU as a whole, approximately 42% of municipal
waste is currently recycled, 24% is incinerated for energy generation and the remaining
34% is consigned to landfill sites. In some of the member states — notably Germany,
Belgium, Austria and the Netherlands — landfill has almost been eliminated through
strong legislation that encourages recycling and energy generation.6 In the United
States, by contrast, only 23% of household waste is recycled, 6.4% is composted, 7.6% is
incinerated to generate energy and 63% goes to landfill.7
The combustible component of MSW may be extracted either chemically by
gasification (to yield gases, liquids and char) or by anaerobic pyrolysis. Alternatively,
anaerobic digestion may be employed to generate methane from solid waste through
the action of specific bacteria. Anaerobic digestion as currently practised requires a wet
waste of relatively high nitrogen content. Consequently, the nitrogen‐to‐carbon ratio of
about 0.03 that is typical of most plant‐based biomass is increased to 0.07 by the addition
of animal manure, sewage sludge or other nitrogen‐rich waste. Anaerobic digesters can
be built at a relatively small scale (a few kW), compared with pyrolysis gasifiers that
normally only become attractive at the MW scale. Such anaerobic digesters are finding
increasing use for generating power in remote areas of countries (e.g., India) that cannot
afford to be connected to an electricity distribution network.
Biogases produced directly from landfill sites, or by the anaerobic digestion of
biomass, contain mixtures of CH4, CO2 and N2, together with various other organic
6 Eurostat 2012 website. http://ec.europa.eu/eurostat
7 Shin, D, 2014, Generation and Disposition of Municipal Solid Waste (MSW) in the United States –A
National Survey, Columbia University Earth Engineering Center.
273
274
Fuel Cell Systems Explained
Table 10.4 Example compositions of biogases.
Biogasa
Source
Agricultural
sludge
Methane (vol.%)
55–65
Ethane (vol.%)
Biogasb
Landfill
gase
Biogasc
Biogasd
Agricultural
sludge
Brewery
effluent
50–70
65–75
57
30–45
30–40
25–35
37
55–70
0
Propane (vol.%)
0
Carbon dioxide (vol.%)
Nitrogen (vol.%)
33–43
2–1
0–2
Small
Hydrogen sulfide (ppm)
<2000
~500
Small
Ammonia (ppm)
<1000
~100
Hydrogen (vol.%)
6
<5000
<1
Small
Higher heating value
MJ (Nm3)−1
23.3
3 −1
Density kg (Nm )
>20
1.16
a)
b)
Paper BP‐12 20th World Gas conference 1997.
Jemsen, J, Tafdrup, S and Chrisensen, J, 1997, Combined utilization of biogas and natural gas,
20th World Gas Conference, 1997, Paper BO‐06,
c) Renewable Energy World, March 1999, p. 75.
d) CADDET, Renewable Energy Newsletter, July 1999, pp. 14–16 (Biogas used in Toshiba 200‐kW
phosphoric acid fuel cell.)
e) CADDET, Renewable Energy Technical Brochure, No. 32, 1996.
compounds. The respective compositions vary widely and, in the case of landfill, depend
on the age of the site. A new site usually produces gas with a high heating value, but this
tends to decrease over a period of time. Some compositions of biogases are listed in
Table 10.4.
Given the relatively high levels of carbon oxides and nitrogen, most biogases have
low heats of combustion and are thus unattractive for use in gas engines or turbines.
Fortunately, this is not a major issue for fuel cells, particularly in the case of molten
carbonate (MCFC) and solid oxide fuel cell (SOFC) systems that are able to handle
very high concentrations of carbon oxides. The same situation holds for phosphoric
acid fuel cells (PAFCs) but to a lesser extent. Many MCFC and PAFC systems fuelled
with landfill gas and/or gas from waste‐water treatment plants have been built and
successfully operated.
Bio‐liquids, such as methanol and ethanol, are acceptable for some fuel‐cell systems,
as described in Sections 6.1 and 6.2, Chapter 6. Methanol can be obtained from syngas
that may be derived from biomass or natural gas, whereas ethanol can be produced
directly by the fermentation of biomass. Alcohols are also attractive because of the
ease by which they can be reformed into hydrogen‐rich gas. This makes the alcohols
suitable for applications such as stationary power backup systems, e.g., for remote
telecommunications towers.
Fuels for Fuel Cells
10.4
Basics of Fuel Processing
10.4.1
Fuel‐Cell Requirements
Fuel processing may be defined as the conversion of the raw primary fuel supplied to a
fuel‐cell system into the fuel gas required by the stack. Each type of stack requires fuel
of a particular quality, as summarized in Table 10.5. Essentially, the lower the operating
temperature of the stack, the more stringent is the level of fuel quality, and therefore the
greater the demand placed on fuel processing. For example, fuel fed to a PAFC stack
needs to be hydrogen‐rich and contains less than about 0.5 vol.% CO, whereas a PEMFC
has to be essentially free of CO. By contrast, both MCFC and SOFC fuel cells are capable
of utilizing this gas internally through the ‘water-gas shift (WGS)’ reaction. Additionally,
unlike PAFCs and PEMFCs, SOFCs and internal‐reforming MCFCs can operate with
methane. It is not widely known, however, that PEMFCs can run directly on some
hydrocarbons such as propane, although the performance is poor.8
Basic explanations of the various technologies for fuel processing are given in the
following sections. Some of the detailed design of individual reactors and systems is
proprietary, of course, but a wealth of information is available from various organizations
who are involved in the development of fuel‐cell systems.
10.4.2
Desulfurization
As noted earlier, natural gas and petroleum liquids contain organic sulfur‐containing
compounds that normally have to be removed before further processing of the given
fuel. Some deactivation of catalysts used in steam reforming can occur with fuel
containing less than 0.2 ppm sulfur, and WGS catalysts are even more intolerant. For
the fuel cell itself, it has been shown that levels of only 1 ppb are sufficient to poison
permanently the anode catalyst in a PEMFC.
Fossil fuels and biofuels often contain a range of sulfur compounds. In the case of
natural gas, sulfur may exist in the form of H2S, and it may be present in the odourant
Table 10.5 Fuel quality for principal types of fuel cell.
Gas species
PEMFC
AFC
PAFC
MCFC
SOFC
H2
Fuel
Fuel
Fuel
Fuel
Fuel
CO
Poison (>10 ppm)
Poison
Poison (>0.5%)
Fuel
a
Fuela
CH4
Diluent
Diluent
Diluent
Diluent
Diluentb
CO2 and H2O
Diluent
Poison
Diluent
Diluent
Diluent
S (as H2S and
COS)
Few studies,
to date
Unknown
Poison (>50
ppm)
Poison
(>0.5 ppm)
Poison
(>1.0 ppm)
a)
b)
b
In reality, CO reacts with H2O producing H2 and CO2 via the shift reaction (see reaction (8.3) in Chapter 8),
and CH4 reacts with H2O to produce H2 and CO faster than it oxidizes as a fuel at the anode.
Methane is a fuel for the internal‐reforming MCFC and SOFC.
8 Baker, BS, 1965, Hydrocarbon Fuel Cell Technology, Academic Press, New York and London.
275
276
Fuel Cell Systems Explained
that has been introduced by the utility company for safety reasons. With petroleum
fractions, the sulfur compounds may be highly aromatic in nature, and gasoline currently
contains some 300–400 ppm of sulfur as organic compounds.
In the drive to reduce emissions from vehicles, regulations have been introduced to limit
the sulfur in both gasoline and diesel fuels. In the United States, for example, the 2004 Tier
2 Gasoline Sulfur Program permitted refiners to produce fuel with a range of sulfur levels,
provided the annual corporate average remained below 30 ppm and no individual batch
exceeded 80 ppm. The current Tier 3 programme lowered the sulfur content of all gasoline
to a maximum of 10 ppm from the beginning of 2017. The Euro 3 standard issued by the
EU in January 2000 set a sulfur limit of 350 ppm for diesel and 150 ppm for gasoline. The
Euro 5 came into effect in 2009 and restricts sulfur in both fuels to 10 ppm. Further
reduction of the allowable levels may require refinement of the methods that are presently
used to desulfurize such fuels — even to the stage of complete removal of sulfur.
There are essentially two methods of desulfurizing fuels. The most common industrial
approach is a process known as hydrodesulfurization (HDS). In the HDS reactor,
any organic sulfur‐containing compounds are converted over a supported nickel–
molybdenum oxide or cobalt–molybdenum oxide catalyst to hydrogen sulfide via
hydrogenolysis reactions such as:
C 2 H 5 2 S 2H 2
2C 2 H6 H2S
(10.1)
The rate of hydrogenolysis increases with temperature. At operating temperatures of
300–400°C and in the presence of excess hydrogen, the reaction essentially goes to
completion. It should also be noted that, at a given temperature, the lighter sulfur
compounds easily undergo hydrogenolysis, whereas the corresponding reaction rate for
odourants such as thiophene (C4H4S) and THT (C4H8S) is much slower. The H2S that is
formed by such reactions is subsequently absorbed by a bed of zinc oxide and therein
converted to zinc sulfide, i.e.,
H2S ZnO
ZnS H2O
(10.2)
The operating conditions and composition of the feed gas determine the choice between
nickel or cobalt catalysts. The optimum temperature for most HDS catalysts lies
between 350 and 400°C, and the catalyst and zinc oxide may be placed in the same
vessel. A popular variation of the traditional industrial HDS process, known as the
PURASPEC™ process, is marketed by Johnson Matthey Process Technologies.
Hydrodesulfurization as a means of removing sulfur to very low levels is ideally suited
to PEMFC or PAFC systems. In such technology, the hydrogen required by reaction (10.1)
is obtained by recycling a small amount of the reformer product, which is rich in hydrogen,
back to the HDS reactor upstream of the reformer. Unfortunately, HDS cannot easily be
applied to internal‐reforming MCFC or SOFC systems, since there is no hydrogen‐rich
stream to feed to the reactor. At least one developer of high‐temperature fuel cells
overcame this problem by including a small reformer reactor into their plant for the sole
purpose of generating hydrogen for HDS.
If HDS is not feasible, then it may be possible to remove sulfur‐containing compounds
from the fuel gas with the aid of an absorbent material. Activated carbon is especially
suitable for use in small systems and can be impregnated with metallic promoters to
enhance the absorption of specific compounds such as hydrogen sulfide. Molecular
Fuels for Fuel Cells
sieves such as the ZSM‐5‐type and faujasite‐type zeolites may also be employed.
The absorption capacity of such materials is, however, quite low, and the beds of
absorbent have to be replaced at regular intervals. These issues may result in a serious
economic disadvantage for large systems.
Some organizations claim to have developed sulfur‐tolerant catalysts for reforming or
POX applications. Argonne National Laboratory, for example, has developed a sulfur‐
tolerant catalyst for their autothermal diesel reformer, but this is only intended to cope
with processing of commercial low‐sulfur diesel fuel. If POX is used to process fuel,
rather than steam reforming, it is likely that the product gas will still contain a few ppm
of sulfur, and its removal will be an essential, but non‐trivial, additional processing step
before feeding the gas to the anode of a PEMFC. Zinc oxide (ZnO) may be employed to
remove traces of H2S that appear in the outlet of a POX reactor, and although the oxide
can be regenerated, it will degrade over time. Consequently, there is a need to devise
desulfurization systems that can operate at moderate temperatures and with high
concentrations of steam in the fuel stream such that they can be used to clean up the
syngas produced by POX. To address this issue, McDermott in the United States
conducted tests on a regenerable zinc oxide bed, and several research groups have also
attempted to improve the thermal durability of highly active ZnO nanoparticles by
supporting them on alumina, silica and carbon. Ceramic monoliths coated with zinc
oxide have also been examined for mobile fuel‐cell systems.
10.4.3
Steam Reforming
Steam reforming is a mature technology that is practised industrially on a large scale for
hydrogen production. Several detailed reviews of the technology have been published,9
and useful data for system design is provided by Twigg.10 The respective basic reforming
reactions for methane and a generic hydrocarbon CnHm are:
CH 4 H2O
C n H m nH 2 O
CO H2O
h f
CO 3H2
nCO
CO2
H2
m
2
h f
206 kJ mol
(10.3)
(10.4)
n H2
41 kJ mol
1
1
(10.5)
The reforming reactions (10.3) and (10.4) and the associated WGS reaction (10.5) are
usually conducted over a supported nickel catalyst at an elevated temperature, typically
above 500°C. Reactions (10.3) and (10.5) are reversible and normally reach equilibrium
over an active catalyst because the rates of reaction are very fast at such high temperatures.
Furthermore, a catalyst that is active for reaction (10.3) nearly always promotes reaction
(10.5). The combination of the two reactions taking place means that the overall product
gas is a mixture of CO, CO2 and H2, together with unconverted CH4 and steam.
9 Rostrup‐Nielsen, JR, 1993, Production of synthesis gas, Catalysis Today, vol. 18, pp. 305–324; also:
Trimm, DR, 2009, Fuel cells — hydrogen production — natural gas conventional steam‐reforming, in
Garche, J, Dyer, CK, Moseley, PT, Ogumi, Z, Rand, DAJ and Scrosati, B (eds.), Encyclopedia of
Electrochemical Power Sources, pp. 203–299, Elsevier, Amsterdam.
10 Twigg, M, 1989, Catalyst Handbook, 2nd edition, Wolfe, London.
277
Fuel Cell Systems Explained
Figure 10.3 Equilibrium
concentrations of products and
reactants for steam reforming of
methane at 100 kPa as a function
of temperature.
60
H2
50
Component / vol.%
278
40
H2O
30
20
CH4
10
CO
CO2
0
500
600
700
Temperature/°C
800
900
The actual composition of the product from the reformer is governed by the temperature
of the reactor, the operating pressure, the composition of the feed gas and the proportion
of the accompanying steam. Graphs and computer models derived from thermodynamic
data are available to determine the composition of the equilibrium product gas under
different operating conditions. By way of example, the composition of the output at
100 kPa is shown in Figure 10.3.
As shown by reaction (10.3), three molecules of H2 and one molecule of CO are
produced for every molecule of CH4 reacted. According to Le Chatelier’s principle,
the equilibrium of this reaction will be moved to the right (i.e., in favour of hydrogen) if
the pressure in the reactor is kept low. Conversely, increasing the pressure during the
course of reforming will favour the formation of CH4, since moving to the left of the
equilibrium reduces the number of molecules in the system. By contrast, the influence of
pressure on the equilibrium position of the WGS reaction (10.5) is very small. Reactions
(10.3) and (10.4) are usually strongly endothermic, and therefore heat needs to be supplied to the reforming reaction to drive it forward to produce H2 and CO. High temperatures (up to 800°C) therefore favour H2 formation, as shown in Figure 10.3.
It is important to note that although the WGS reaction (10.5) occurs simultaneously
with the steam reforming reaction on most catalysts, at the high temperatures that are
necessary for hydrogen generation, the equilibrium point for the reaction is well to the
left of the equation. Accordingly, by no means all the CO will be converted to CO2.
Consequently, processing will be required for fuel‐cell systems that require low levels of
CO and is examined further in Section 10.4.11.
Remarkably, steam reforming is not always endothermic. For example, in the case of a
petroleum hydrocarbon such as naphtha, with the empirical formula CH2.2, the reaction
is most endothermic when the hydrocarbon reacts with steam to give only oxides of
carbon and hydrogen. Steam reforming of naphtha is therefore most endothermic when
carried out at high temperatures. It is less endothermic and eventually exothermic
(liberates heat) as the temperature is lowered because this favours the reverse of reaction
(10.3), i.e., the formation of CH4. The fact that, depending on the temperature and
pressure of the reactor, the steam reforming of naphtha can be either endothermic or
exothermic is demonstrated by the reactions listed in Table 10.6.
Fuels for Fuel Cells
Table 10.6 Typical heats of reaction in naphtha reforming at different temperatures, pressure
and steam‐to‐carbon ratios.
Pressure
(kPa)
Temperature
(°C)
Steam‐
to‐carbon
ratioa
∆H (25°C)
(kJ mol−1
CH2.2)
2 070
800
3.0
CH2.2 3H2O
0.2CH 4 0.4 CO2 0.4 CO 1.94 H2 1.81H2O
+102.5
2 760
750
3.0
CH2.2 3H2O
0.35CH 4 0.4 CO2 0.25CO 1.5H2 1.95H2O
+75.0
31 050
450
2.0
CH2.2 2H2O
0.75CH 4 0.25CO2 0.14 H2 1.5H2O
–48.0
Reaction
Source: Kramer, GJ, Wieldraaijer, W, Biesheuvel, PM & Kuipers, HCPE, 2001, The determining factor
for catalysts selectivity in Shell’s catalytic partial oxidation process, American Chemical Society,
Fuel Chemistry Division Preprints, vol. 46(2), pp. 659–660.
a) Ratio of the number of moles of steam to the number of moles of carbon in the steam + fuel fed to the
reactor.
In summary, steam reforming has the following implications for fuel‐cell systems.
The reforming of natural gas will invariably be endothermic, and heat will have to be
supplied to the reformer at a sufficiently high temperature to ensure a reasonable degree
of conversion. Naphtha and similar petroleum fractions (e.g., gasoline, diesel) will also
react under endothermic conditions when hydrogen is the preferred product and the
operating temperature is kept high. If, however, steam reforming of naphtha is carried
out at more moderate temperatures (up to 600°C), the product will contain a significant
concentration of CH4 and the reaction will be less endothermic.
‘Dry reforming’ (also known as ‘CO2 reforming’) can be carried out if there is no ready
source of steam as follows:
CH 4 CO2
2CO 2H2
h f
247 kJ mol
1
(10.6)
This reaction may occur in internal‐reforming fuel cells, for example, the MCFC, when
anode exhaust gas, which contains CO2 and water, is recycled to the inlet of the fuel cell.
‘Mixed reforming’ is a term that is sometimes used to describe a hybrid approach where
both steam and CO2 are used to reform the fuel. Dry and mixed reforming have
energy and environmental advantages compared with traditional steam reforming. The
reactions are catalysed by nickel, but deactivation due to carbon formation and nickel
sintering can be particularly severe.
Hydrocarbons such as methane, light distillates and naphtha are not the only fuels
suitable for steam reforming. Alcohols will also react with steam to produce hydrogen
and carbon oxides. Methanol, for example, yields:
CH3OH H2O
3H2 CO2
h f
49.7 kJ mol
1
(10.7)
The fact that this reaction is mildly endothermic is one of the reasons why methanol
has been favoured by some vehicle manufacturers as a possible fuel for FCVs.
279
280
Fuel Cell Systems Explained
In addition, methanol is a suitable fuel for remote stationary power systems. Little heat
needs to be supplied both to overcome the inherent heat losses and to sustain the
methanol reforming reaction, which will readily occur at modest temperatures (e.g.,
250–300°C) over catalysts of mild activity such as copper supported on zinc oxide.
Although CO does not feature as a principal product of reaction (10.7), this does not
mean that it will not be present. The WGS reaction (10.5) is reversible, and with an
active catalyst, even at moderate temperatures, some CO will be produced from CO2
via the reverse shift reaction. Nonetheless, the CO level in reformed methanol will be
acceptable for PAFCs and high‐temperature PEMFCs. For low‐temperature PEMFCs,
the CO content has to be lowered using one of the methods described in Section 10.4.11.
10.4.4
Carbon Formation and Pre‐Reforming
Carbon formation by decomposition of the fuel gas is one of the most critical problems
that may be encountered during the operation of fuel‐cell systems. This reaction can
take place in several areas of the system where hot fuel gas is present. Natural gas, for
example, will decompose when heated in the absence of air or steam at temperatures
above about 650°C via pyrolysis reactions such as:
CH 4
C 2H 2
h f
75 kJ mol
1
(10.8)
Higher hydrocarbons tend to decompose more easily than methane, and therefore the
likelihood of carbon formation is greater with vapourized liquid petroleum fuels than
with natural gas.
Another source of carbon formation is from the disproportionation of CO via the
Boudouard reaction, i.e.,
2CO
C CO2
(10.9)
The reaction is catalysed by metals such as nickel, and consequently there is a high
probability that it will occur on steam reforming catalysts that contain nickel and on the
walls of stainless‐steel reactors. Fortunately, there is a simple expedient to reduce the
degree of carbon formation via reactions (10.8) and (10.9), namely, to add excess steam
to the fuel stream. The principal effect of this is to promote the WGS reaction (10.5),
which results in a reduction of the partial pressure of CO in the fuel gas stream.
Moreover, steam encourages the carbon gasification reaction, which is also very fast, as
expressed by:
C H2 O
CO H2
(10.10)
By considering the thermodynamics of the system, it is possible to calculate the
minimum amount of steam that has to be added to a hydrocarbon fuel gas to avoid carbon
deposition. The procedure is based on the assumption that a given mixture of fuel gas and
steam interacts through reactions (10.3), (10.4) and (10.5) to produce a gas that is in
thermodynamic equilibrium with respect to reactions (10.3) and (10.5) at the particular
temperature and pressure of operation. The measured or observed partial pressures of
CO and CO2 are used to calculate an equilibrium constant for the Boudouard reaction
(10.9). If, for the observed temperature, the calculated constant is greater than the
theoretical value, then carbon deposition is expected on thermodynamic grounds.
Fuels for Fuel Cells
If, however, the calculated constant is lower than theory predicts, then the gas is said to
be in a safe region and carbon deposition will not occur. In practice, a steam‐to‐carbon
ratio of 2.0–3.0 is normally employed in steam reforming systems so that carbon
deposition may be avoided with an adequate margin of safety.
A particular type of carbon formation that can occur on metals is known as
‘carburization’ and can lead to spalling of metal, which is referred to as ‘metal dusting’.
Again, it is important to reduce the possibility of this phenomenon developing in fuel‐
cell systems, and some developers have used copper‐coated stainless steel in their fuel
gas preheaters to minimize its occurrence.
Carbon formation on steam reforming catalysts has been the subject of intense study
and is well understood. Carbon produced via the pyrolysis reaction (10.8) and the
Boudouard reaction (10.9) adopts different forms, of which the most damaging are
filaments that appear to ‘grow’ on the nickel crystallites within the catalyst. Such carbon
filaments can develop very rapidly, for example, if the steam supply to the reformer
reactor is cut off suddenly. In such an event, the consequences can be disastrous with
carbon formation occurring within seconds, thereby causing permanent breakdown of
the catalyst that, in turn, leads to plugging of the reactor. This hazard emphasizes the
importance of process control during the operation of fuel‐cell systems, especially if the
steam is obtained by recycling product from the stack. Commercial steam reforming
catalysts contain elements such as potassium and molybdenum that are known to
inhibit carbon formation on the catalyst surface.
The propensity for carbon formation can also be diminished by ‘pre‐reforming’ the fuel
gas before it is fed to the reformer reactor. Pre‐reforming is a term commonly used in
industry to describe the preferential conversion of high molecular weight hydrocarbons to
hydrogen and carbon oxides via steam reforming at relatively low temperatures (typically
250–500°C). This process step is carried out in an adiabatic reactor, i.e., one that has
neither external heat nor cooling applied. The gas from the exit of a pre‐reformer therefore
consists mainly of methane and steam, together with small amounts of hydrogen and
carbon oxides; the exact composition depends on the temperature of the reactor.
A pre‐reformer arrangement that was designed for the demonstration of an internal‐
reforming SOFC stack is shown in Figure 10.4. The facility was built not only to reform
the higher hydrocarbons from the natural gas but also to convert some 15% of the CH4
and thereby provide sufficient hydrogen at the stack inlet to maintain the SOFC anodes
in a reduced condition. The pre‐reforming also had the beneficial effect of reducing
thermal stresses within the stack since only 85% of CH4 was reformed internally.
10.4.5
Internal Reforming
In the foregoing discussion of fuel processing, it has been assumed that steam reforming
is conducted in one or more reactors that are external to the fuel‐cell stack — a practice
therefore known as ‘external reforming’. For many years, developers have been aware that
the heat to sustain the endothermic reforming of low molecular weight hydrocarbons
(e.g., natural gas) can be provided by the electrochemical reaction in the stack. This
feature has led to various elegant concepts of internal reforming that have been applied
to MCFCs and SOFCs, on account of their high operating temperatures. The option is
not available for low‐temperature PEMFC, alkaline fuel cell (AFC) and PAFC systems.
In contrast to the steam reforming reactions (10.3) and (10.4), the fuel‐cell
reactions are always exothermic, mainly due to heat production in the cell caused by
281
282
Fuel Cell Systems Explained
490 °C
Natural gas
2.74 kW
520 °C
15 °C
Anode exhaust burner
and steam recovery
subsystem
Catalytic
adiabatic
pre-reformer
Desulfurizer
(activated carbon)
371 °C
618 °C
850 °C
SOFC
anodes
5.23 kW
700 °C
Figure 10.4 Pre‐reformer system devised for a Siemens 50‐kW SOFC demonstration.
internal resistances. Under practical conditions, with a cell voltage of 0.78 V, the
heat liberated in a fuel cell amounts to 470 kJ per mol CH4. By carrying out steam
reforming of CH4 within the MCFC or SOFC stack, about half of the heat liberated
by the fuel‐cell reactions can therefore be utilized by the steam reforming reaction
(10.3). This obviously reduces significantly the need for stack cooling, which is
normally achieved by flowing excess air through the cathode. The lower airflow
required by internal‐reforming stacks significantly improves the electrical efficiency
of the overall system.
Developers of internal‐reforming fuel cells have generally adopted one of two
approaches. These are usually referred to as direct internal reforming (DIR) and indirect
internal reforming (IIR) and are illustrated schematically in Figure 10.5. In some cases,
a combination of both approaches has been adopted.
The application of internal reforming offers several additional advantages compared
with external reforming, as follows:
●
●
●
●
●
System cost is reduced because a separate external reactor is not required.
Less steam is necessary with DIR since the anode reaction in both the SOFC and the
MCFC produces steam.
The generation of hydrogen over the anode of a DIR cell may lead to a more even
distribution of temperature throughout the cell.
The methane conversion is high, especially in DIR systems where the cell consumes
hydrogen as it is produced.
The efficiency of the system is high because internal reforming can provide an elegant
method of cooling the stack and thereby decrease the need for excess air at the cathode.
This advantage, in turn, lowers the requirement for air compression and recirculation.
Fuels for Fuel Cells
Reforming
catalyst
CH4 + steam
Anode
exhaust
CO2 + steam
Indirect
reforming
Direct
reforming
Heat
H2 and
CO-rich gas
Anode
CO3=
Electrolyte
Cathode
Air + CO2
Cathode
exhaust
Figure 10.5 Schematic representation of direct and indirect internal reforming.
10.4.5.1
Indirect Internal Reforming (IIR)
Also known as ‘integrated reforming’, IIR involves conversion of CH4 in reformer
reactors that are positioned in close thermal contact with the stack. Thus the reforming
reaction and electrochemical reactions are separated. An example of this type of
arrangement alternates plate reformers with groups of five or six cells. The reformate
from each plate is fed to a neighbouring group of fuel cells. Although the high rate of
heat transfer between cells and reformer plates provides a benefit, IIR systems suffer
from the fact that heat is transferred effectively only from cells most adjacent to the
reformers and also that steam for reforming must be raised separately. A variation in
design places the reforming catalyst in the gas‐distribution path of each cell.
10.4.5.2
Direct Internal Reforming (DIR)
In the DIR approach, the reforming reactions are performed in close proximity to the
anode of each cell within the stack, as shown in Figure 10.5. In the case of the MCFC,
reforming catalysts are placed within the gas flow channels of the anode. For the SOFC,
the higher temperature of operation and the fact that the anode usually has a high
nickel content and high surface area mean that steam reforming can occur directly on
the anode, and therefore a reforming catalyst may not actually be required. Direct
internal reforming has the distinct advantage of not only affording good heat transfer
between the fuel cell and the reforming reaction but also ensuring chemical
integration, that is, product steam from the anode reaction is used directly by the
reforming reaction.
Although endothermic reforming absorbs some of the heat produced by MCFC or
SOFC stacks with internal reforming, it is not sufficient to remove completely the need
to cool the stack. Indeed, the combination of endothermic and exothermic reactions
that occur in a stack with internal reforming can promote large temperature gradients
within the cell and stack hardware. Such behaviour can arise because steam reforming
is both fast and endothermic and thus cause a sharp fall in temperature close to the
283
284
Fuel Cell Systems Explained
anode inlet. By contrast, the exothermic fuel‐cell reactions lead to a rise in temperature
towards the anode outlet. To minimize such variations in temperature within an internal‐
reforming MCFC or SOFC stack, a counter‐flow configuration of fuel and oxidant is
usually employed rather than co‐flow.
Finally, it should be understood that internal reforming may, in principle, be applied
to several hydrocarbon fuels that include natural gas and vapourized liquids such as
naphtha and kerosenes. Coal gases are particularly attractive for internal‐reforming
MCFC and SOFC stacks, since not only are CO and H2 consumed directly as fuels, but
residual CH4 (as may be present in the product from the BG–Lurgi slagging gasifier; see
Table 10.2) is also internally reformed.
10.4.6
Direct Hydrocarbon Oxidation
Direct hydrocarbon oxidation refers to the conversion of a hydrocarbon fuel directly to
steam and CO2 without first undergoing conversion to hydrogen by steam reforming
or POX. The Gibbs free energy change for the direct oxidation reaction of CH4 to CO2 and
steam is −796.5 kJ mol−1, which is very close to the change in enthalpy (∆h = −802.5 kJ mol−1).
In other words, if CH4 could be oxidized directly, most of the heat of reaction would be
converted directly into electricity, with a maximum efficiency of
g
h
796.5
100 99.2%
802.5
(10.11)
The high efficiency should be qualified because H2O and CO2 produced by the direct
oxidation of methane can react with fresh methane via the reforming reaction over the
nickel catalyst in an MCFC or the nickel‐containing anode in the SOFC. The only way
therefore to establish whether direct oxidation is actually taking place in a fuel cell is to use
an anode catalyst that does not promote the steam reforming reaction. The first
demonstration of the feasibility of direct methane oxidation was provided by replacing
the nickel cermet that was normally employed as the anode in SOFCs with a copper
cermet.11 The copper cermet also addressed the issue of carbon formation. As discussed
in Section 10.4.4, carbon has a propensity to deposit on nickel‐containing materials at
temperatures even as low as 600°C. With a copper‐cermet anode, however, carbon
formation is avoided and direct oxidation can occur. Furthermore, it has been found
that copper cermets are stable in hydrocarbon environments as long as the temperature
is limited.
Apart from the increased conversion efficiency afforded by direct oxidation, a further
benefit is that steam does not need be supplied with the fuel and thereby leads to systems
of much simpler design. The benefits of high conversion efficiency and lower steam
requirement and resistance to carbon fouling have encouraged various investigations of
novel ceramic anodes for SOFCs made of mixed ionic–electronic conductor (MIEC)
materials, as has been discussed in Section 9.2.3.3, Chapter 9.
11 Park, S, Vohs, JM and Gorte, RJ, 2000, Direct oxidation of hydrocarbons in a solid‐oxide fuel cell,
Nature, vol. 404, pp. 265–267.
Fuels for Fuel Cells
It is also known that direct hydrocarbon oxidation can occur in fuel cells that use
aqueous acid electrolyte. For instance, in the 1960s, propane was found to decompose on
platinum catalysts at moderate temperatures (below 200°C) to form protons, electrons
and CO2 (via reaction with water from the aqueous acid electrolyte). As in both PEMFC
and PAFC systems, the protons migrate to the cathode where they are oxidized with air
and electrons to form water. Researchers may find it worthwhile to revisit some of this
earlier work to examine whether it holds the key to new ways of fuelling fuel cells.
10.4.7
Partial Oxidation and Autothermal Reforming
As an alternative to steam reforming, methane and other hydrocarbons may be
converted to syngas and hence hydrogen for fuel cells by ‘partial oxidation’, i.e.,
CH 4
1
O
2 2
CO 2H2
h f
247 kJ mol
1
(10.12)
Partial oxidation can be carried out at high temperatures (typically 1200–1500°C) in the
gas phase without a catalyst. The process has the advantage over catalytic processes in
that it is not necessary to remove materials such as sulfur compounds, although sulfur
would have to be extracted from the product at a later stage (as sulfur dioxide, SO2) if it
is to be fed to a fuel‐cell stack. High‐temperature POX can also handle much heavier
petroleum fractions than catalytic processes and is therefore attractive for processing
diesel, other liquid fuels and residual oil fractions. Gas‐phase POX has been performed
in large facilities by several companies, but it does not scale down well, and management
of the reaction is problematic.
A greater level of control of the POX reaction can be achieved by lowering the
temperature and employing a catalyst. The process is then termed ‘catalytic partial
oxidation’ (CPO). Catalysts for CPO tend to be either a platinum‐group metal or nickel
supported on a ceramic oxide. At temperatures of around 1000°C, the CPO reaction is very
fast, and an industrial version of the process developed by Eni S.p.A is known as ‘short
contact time catalytic partial oxidation (SCT‐CPO)’. The CPO reaction occurs within a
few milliseconds in a high‐temperature and thin (<1 µm) solid–gas interphase zone that
surrounds the catalyst particles and does not propagate into the gas phase. These
conditions favour the formation of the primary reaction products (H2 and CO) and allow
the use of several hydrocarbon feedstocks that may also contain sulfur or aromatic
compounds. Catalytic partial oxidation has been the subject of much research over the past
10–15 years with Haldor Topsøe A/S and Eni S.p.A establishing the technology for industrial
production of hydrogen, gas‐to‐liquid (GTL) processes and refinery operations.
It should be noted that the POX reaction (10.12) produces less H2 per molecule of
CH4 than the steam reforming reaction (10.3). Accordingly, POX (either gas phase or
CPO) is usually less efficient than steam reforming for fuel‐cell applications. Reaction
(10.12) is effectively the summation of the steam reforming and oxidation reactions; it
can be considered that about half of the fuel that is converted into hydrogen is oxidized
to provide heat for steam reforming. Heat from the fuel cell is not utilized by POX or
CPO reactors, and the net effect is a reduction in overall system efficiency. Another
disadvantage of POX occurs when air is used as the oxidant. The presence of nitrogen
from the air reduces the partial pressure of hydrogen at the fuel cell that, in turn, lowers
the cell voltage (as determined by the Nernst equation) so that again there is a decrease
285
286
Fuel Cell Systems Explained
in system efficiency. To offset these negative aspects, the key advantages of POX with
air are that steam is not required and the process can be achieved in a much smaller
reactor than that used for steam reforming. Consequently, POX may be considered for
applications where system simplicity is regarded as more important than high electrical
conversion efficiency as, for example, in small‐scale cogeneration systems.
If the steam reforming and CPO reactions are carried out in the same reactor and
over the same catalyst, the process is generally known as ‘autothermal reforming’. It
should be pointed out, however, that the terms ‘autothermal reforming’ and ‘partial
oxidation’ are often loosely employed in the literature and care needs to be exercised
when comparing data from reports of different fuel‐processing systems. Autothermal
reforming involves feeding fuel, steam and oxygen (or air) simultaneously to the reactor.
By adjusting the flows of these reactants, as well as allowing for heat transferred out of
the reactor in the product gas and any inevitable heat losses to the environment, a
thermoneutral condition can be sustained in which the heat generated by POX exactly
matches the heat consumed by steam reforming.
In terms of reaction mechanisms, several studies have determined the relative rates of
the steam reforming and the CPO reactions when conducted simultaneously over
different catalysts. From such work, it has been concluded that in many cases the CPO
reaction is brought to equilibrium much faster than the steam reforming reaction. This
sequence, which has been termed the ‘indirect mechanism’ of POX, is promoted by
supported nickel catalysts. By contrast, supported ruthenium and rhodium catalysts
activate an alternative ‘direct mechanism’ whereby oxidation, steam reforming and shift
reactions occur in parallel.
In some processes that have been described as CPO, both steam and oxidant are fed with
the fuel. The Shell POX process is one example of such practice (Kramer et al., 2001).12 The
process uses a proprietary vertical tubular reactor design that contains a bed of platinum‐
group catalyst. The partial oxidation reaction occurs at the top of the bed where the rate is
limited by mass transfer of the reactants. Further down the bed, the steam reforming and
WGS reactions bring the gas to equilibrium. In other examples of CPO where the steam
oxidation and reforming reactions operate in parallel, the rates of the reactions are not
limited by mass transfer and are brought to equilibrium without gain or loss of heat.
The advantages of autothermal reforming and CPO are that less steam is required
compared with conventional steam reforming and that all of the heat for the reforming
reaction is provided by partial combustion of the fuel. Both approaches are attractive
for application with PEMFCs and AFCs as these cells do generate heat at a temperature
that is sufficiently high for the steam reforming of fuel.
10.4.8
Solar–Thermal Reforming
The heat for steam reforming of hydrocarbons such as methane can, in principle, be
obtained directly from the sun. As the resulting syngas would contain a substantial
amount of embodied solar energy (up to 25%), solar–thermal reforming offers the
12 Kramer, GJ, Wieldraaijer, W, Biesheuvel, PM and Kuipers, HCPE, 2001, The determining factor for
catalysts selectivity in Shell’s catalytic partial oxidation process, American Chemical Society, Fuel Chemistry
Division Preprints, vol. 46(2), pp. 659–660.
Fuels for Fuel Cells
prospects of high thermal efficiencies and greatly reduced emissions of carbon dioxide.
Moreover, the emissions would be in concentrated form and thus more amenable to gas
separation. Over the past 20 years, a considerable amount of work has been undertaken
on solar–thermal reforming in several countries, and broadly three types of solar
receiver–reactor have been proposed, 13 namely:
1) Molten sodium, reflux heat‐pipe receiver–reactor.
2) Directly irradiated, volumetric receiver–reactor (DIVRR).
3) Cavity receiver with conventional, tubular catalytic reactors.
In each design, solar energy is directly focused on the receiver–reactor or transmitted
via the use of heat‐transfer fluid from the solar receiver to the reactor. To achieve the
required high temperature, three common methods are employed:
1) A simple parabolic dish that focuses the sun’s rays on to a thermal receiver mounted
above the dish at its focal point.
2) An array of thousands of individual mirrors (‘heliostats’) around a central receiver
set on top of a tall tower.
3) Parabolic trough mirrors that track the sun as it crosses the sky and that have
receivers located at their foci.
The three methods are illustrated in Figure 10.6.
Solar–thermal reforming has a number of challenges. The first is that a large area of
land has to be available to accommodate the dish, troughs or heliostats. Another concern
is that solar energy is highly variable through the day and is not available at night.
Therefore, some means of energy or heat storage has to be provided to ensure continuous
production of hydrogen. There are also issues with the reforming catalyst in that
temperatures above 850°C are required with conventional reforming catalysts to obtain
an adequate yield of hydrogen, and both system performance and catalyst integrity are
compromised if the temperature is allowed to fall and fluctuate excessively. The situation
can be improved by employing catalysts that incorporate more active precious metals,
e.g., ruthenium or rhodium, or by carrying out the reforming in a membrane reactor to
achieve direct separation of hydrogen.
10.4.9
Sorbent‐Enhanced Reforming
By mixing a solid absorbent for CO2 (e.g., calcined dolomite, CaO) with the reforming
catalyst, it is possible to combine both the steam reforming and WGS reactions into a
single step and, at the same time, reduce the temperature of the former from about 900
to 400–500°C. Removal of CO2 from the reaction zone influences the equilibrium of
the combined reaction so that the production of hydrogen is enhanced, while CO is
oxidized to CO2. A typical product gas from the sorbent‐enhanced reforming of CH4
is composed of 90 vol.% H2 and about 10 vol.% unreacted CH4, with a small percentage
of CO2 and a trace of CO.
13 Stein, W, Edwards, J, Hinkley, J and Sattler, C, 2012, Natural gas: solar‐thermal steam‐reforming, in
Garche, J, Dyer, CK, Moseley, PT, Ogumi, Z, Rand, DAJ and Scrosati, B (eds.), Encyclopedia of
Electrochemical Power Sources, pp. 300–312, Elsevier, Amsterdam.
287
288
Fuel Cell Systems Explained
(a)
Receiver or
engine/receiver
Concentrator
reflective surface
(b)
Receiver
Heliostats
Tower
(c)
Concentrator
reflective surface
Receiver
Tracking
mechanism
Figure 10.6 Three main designs of solar furnace: (a) parabolic dish, (b) central receiver and
(c) parabolic trough.
When the sorbent has become saturated with CO2, it is regenerated by purging with
steam. The exit stream is condensed and the released CO2 can be captured ready for
compression and conveyance to underground storage. For continuous operation, two
parallel reactors are required — one bed undergoes reaction, while the other is
regenerated. The sorbent‐enhanced process offers an elegant approach in that it
removes the need for a second shift and for additional gas‐separation steps. Also, the
lower temperature of operation results in reduced heat loss and permits the use of
cheaper materials of construction. Sorbent‐enhanced reforming has been the subject
of research in many countries for the past twenty or so years, but there are technical
challenges to be overcome before it can become a commercial proposition. These
include the long‐term durability of the sorbent and catalyst and the scale‐up and
switching of reactant streams to the two reactors.
Fuels for Fuel Cells
10.4.10
Hydrogen Generation by Pyrolysis or Thermal Cracking of Hydrocarbons
An alternative to all of the aforementioned methods of generating hydrogen from
hydrocarbons is simply to heat the fuel in the absence of air, a process commonly referred
to as ‘pyrolysis’. The hydrocarbon ‘cracks’ or decomposes into hydrogen and solid carbon
(soot). The process is ideally suited to simple hydrocarbon fuels; otherwise various by‐
products may be formed (e.g., acetylene and other alkenes). The advantage of thermal
cracking is that the hydrogen produced can be very pure and, if no catalyst is present, the
carbon can be separated as a solid, which is easier to handle than gaseous CO2. On the
other hand, if pyrolysis is enhanced with a catalyst, then extraction of carbon becomes a
challenge. In principle, removal can be achieved by shutting off the supply of fuel and
admitting air to the reactor to burn off the carbon as CO2. Switching the flow of fuel and
oxidant is simple in theory, but there are real difficulties, not least of which are the safety
implications of admitting fuel and air into a reactor at high temperatures. Control of the
pyrolysis is critical; otherwise too much carbon will accumulate and consequently will
irrevocably damage any catalyst that may be employed. Excessive carbon may also be
formed in the absence of a catalyst — indeed, to such an extent that the reactor becomes
plugged and accordingly creates difficulty with the removal of the solid and/or with
establishing a sufficient flow of oxidizing gas to burn off the deposited material. Despite
these substantial problems, pyrolysis is being considered seriously as an option for some
fuel‐cell systems. For instance, thermal cracking of propane has been proposed as a
source of hydrogen for small PEMFC systems.14
The carbon build‐up problem with thermal pyrolysis can be avoided by the application
of thermal plasma technology (see Box 10.2). This is because a thermal plasma, which is
created by an electric arc, is characterized by temperatures of 3000–10 000°C, and under
such conditions no catalyst is required. A further advantage is that the reactions are
very fast and therefore allow a much more compact and lighter design than that for a
conventional catalytic processor. By adding steam and oxygen, a plasma reactor can
operate as a reformer and produce a syngas. Non‐thermal plasmas (notably the sliding
discharge reactor; see Box 10.2) have also been investigated for reforming fuels.15
A compelling feature of plasmas is their ability to break down high molecular weight
hydrocarbons. In most cases, however, control of the plasma is both difficult and energy
demanding and thereby results in poor conversion and variable product composition.
Recently, researchers have sought to overcome these limitations by combining a reforming
catalyst either within or external to the plasma reactor. Although still at the laboratory
stage, such approaches may lead to an energy‐efficient device that fulfils the needs of a
small‐scale hydrogen generator for fuel‐cell systems.16
14 Wang, Y, Shah, N and Huffman, GP, 2003, Production of pure hydrogen and novel carbon nanotube
structures by catalytic decomposition of propane and cyclohexane, Prepr. Pap.‐Am. Chem. Soc., Div. Fuel Chem.,
vol. 48(2), p. 901.
15 Paulmier, T and Fulcheri, L, 2005, Use of non‐thermal plasma for hydrocarbon reforming, Chemical
Engineering Journal, vol. 6, pp. 59–71.
16 Tu, X and Whitehead, C, 2014, Plasma dry reforming of methane in an atmospheric pressure AC gliding
arc discharge: Co‐generation of syngas and carbon nanomaterial, International Journal of Hydrogen Energy,
vol. 39(18), pp. 9658–9669.
289
290
Fuel Cell Systems Explained
Box 10.2 What is a plasma?
A plasma is a partially or fully ionized gas composed of positive ions and free electrons in
proportions that result in more or less no overall electric charge. Based on the relative
temperature of the electrons, ions and neutral particles, plasmas are classified as ‘thermal’
or ‘non‐thermal’. Thermal plasmas have the electrons and heavier particles at the same
temperature (i.e., they are in thermal equilibrium with each other), whereas non‐thermal
plasmas have the ions and neutrals at much lower temperatures (e.g., room temperature)
and the electrons are much ‘hotter’. A thermal plasma is formed when an arc is discharged
between two electrodes. Examples of non‐thermal plasmas are as follows.
●
●
●
●
●
Glow discharges are produced when a high DC voltage is applied between metal
electrodes in a reactor in which gas is at a low pressure (0.01–1.0 kPa) — the familiar
neon and fluorescent light tubes are good examples. A glow discharge can also be
formed with an alternating current at radio frequencies.
Silent discharges, also known as ‘corona’ discharges, are obtained in gases at low or
moderate pressures in which the discharge is struck between conductors with rounded
or pointed tips.
Dielectric barrier discharge is created when a high voltage is applied between two
electrodes that have non‐conductive coatings to prevent an arc from developing.
Capacitative discharges are generated when radio‐frequency (RF) power is applied to
one electrode with a grounded electrode held at a small distance of separation,
typically 1 cm.
Electrodeless or microwave discharges are produced by an electric field from a coil
wrapped around the reactor.
10.4.11
Further Fuel Processing: Removal of Carbon Monoxide
A steam reformer running on natural gas and operating at atmospheric pressure with
an outlet temperature of 800°C produces a gas that consists of some 75 vol.% H2, 15 vol.%
CO and 10 vol.% CO2 on a dry basis. It is necessary to lower the concentration of CO in
such a gas before it can be fed to a PAFC or a PEMFC.
As noted earlier, the WGS reaction (10.5) takes place on active catalysts at the same
time as the basic steam reforming reaction (10.4). If both reactions are at thermodynamic
equilibrium, high temperatures favour the production of CO, i.e., reaction (10.5) is
shifted to the left. The first approach to reducing the CO content of a reformed fuel is
therefore to cool the product gas from the steam reformer and pass it through a reactor
with a catalyst that only promotes the WGS reaction. This has the effect of converting
CO into CO2. Depending on the composition of the product gas from the reformer,
more than one shift reactor may be necessary to lower the CO to an acceptable level. An
iron–chromium catalyst is found to be effective for promoting the WGS reaction at
relatively high temperatures (400–500°C), and this may be followed by further cooling
of the gas before passing to a second low‐temperature reactor (200–250°C) with a
copper catalyst. At this lower temperature, the proportion of CO exiting the reactor
will typically be about 0.25–0.5 vol.%, and so these two stages of shift conversion are
sufficient to decrease the CO content to meet the needs of the PAFC. On the other
Fuels for Fuel Cells
hand, the level is equivalent to 2500–5000 ppm and exceeds the limit for typical PEMFCs
by two orders of magnitude. It is similar to the level of CO in the product from a methanol
reformer, and therefore further fuel processing is required for a PEMFC system.
Until recently, the WGS reactors in fuel‐cell systems at the kW-scale have utilized
industrial iron and copper catalysts. These are easily poisoned by sulfur and also present
a safety hazard as the catalyst in the reduced state becomes pyrophoric when exposed
to air. In addition, the reactors are large in comparison with the stacks, the reformer
reactor and other balance‐of‐plant components. To improve the situation, some
developers of fuel cells are working on novel catalysts that are able to operate with
higher space velocities (i.e., are more active) and at low temperatures. The Selectra™
Shift catalyst supplied by Engelhard is a non‐pyrophoric base metal material that is
an alternative to the traditional low‐temperature ZnO catalyst. Cobalt oxide and
molybdenum oxide supported on alumina has also been used as a low‐cost, low‐
temperature WGS catalyst in a sulfided form. This particular composition is insensitive
to sulfur, a feature that makes it even more attractive in comparison with the traditional
zinc–copper materials.
Precious metal catalysts, e.g., platinum on ceria (CeO2) developed by Nextech in the
United States, platinum on mixed cerium–lanthanum oxides and gold on ceria, are
non‐pyrophoric, are tolerant to sulfur and may become viable options, provided costs
can be kept reasonably low. Some transition metal carbides also provide sulfur tolerance
and relatively high WGS activity. Nevertheless, there is clearly room for improvement
in WGS catalysts for fuel‐cell systems, and consequently research in this area is quite
active at the moment. In the end, there will probably be a trade‐off between cost and
performance, as well as for the other catalytic steps in fuel processing.
For the PEMFC, further CO removal is essential after the WGS reactors. This is
usually carried out by one of the following three methods:
1) Preferential oxidation (PROX) in a reactor to which a small amount of air (around
2 vol.%) is added to the fuel stream, which then passes over a precious metal catalyst.
Typical catalysts include Pt–Al2O3, Ru–Al2O3, Rh–Al2O3, Au–MnOx, Pt‐Ru–Al2O3
and Ir‐based materials such as 5 wt.% Ir–(CoOx–Al2O3)–carbon. The catalyst
preferentially absorbs the CO, rather than the H2, and reacts with the oxygen in the
air to produce CO2. As well as the obvious issue of the cost of the precious metal
catalysts, the PROX reaction is exothermic, and therefore the reactor may require
cooling so that temperature control can be problematic. Since there is the presence
of H2, CO and O2 at an elevated temperature with a noble metal catalyst, measures
must be taken to ensure that an explosive mixture is not produced. This eventuality
could be a particular concern where the flow rate of the gas is highly variable, such
as with a PEMFC system on board a vehicle.
2) Methanation of the CO is an approach that reduces the danger of producing
explosive gas mixtures. The reaction is the opposite of the steam reforming of CH4,
i.e., reaction (10.3), namely,
CO 3H2
CH 4 H2O
hf
206 kJ mol
1
(10.13)
The reaction is conducted in a small catalytic reactor placed close to the fuel‐cell
inlet. Methanation has the obvious disadvantage that it consumes some hydrogen,
which will reduce the electrical efficiency of the fuel‐cell system by a small amount.
291
292
Fuel Cell Systems Explained
The methane produced by reaction (10.13) does not poison the PEMFC catalyst but
simply dilutes the fuel within the stack. Such minor effects can be tolerated in a
PEMFC system in which a simple methanation catalyst operating at about 200°C can
reduce the level of CO to less than 10 ppm. In the case of reformed methanol, the
methanation catalyst will also ensure that any unconverted methanol from the
reformer reactor is converted variously to CH4, H2 and CO2.
3) Palladium or platinum membranes can be used to separate and purify the H2. This
is mature technology that has been used for many years to produce H2 of exceptional
purity, despite the fact that such membranes are expensive. Further discussion of
membranes is given in Section 10.5.
Workers at the National Renewable Energy Laboratory in the United States have been
attempting to improve the WGS reaction through the use of bacteria, and this research
has been taken up by other groups in recent years. If this method proves to be successful,
there may be less need for the final CO clean‐up stage. To date, however, the reaction
rates for the biological procedure have been two orders of magnitude below those
experienced with the traditional catalytic systems. Further discussion of biological
systems is given in Section 10.10.
Electrochemical oxidation is a further and very different approach to CO removal for
the PEMFC. Two methods have been investigated, namely, (i) oxidation in a reactor
placed before the fuel cell17,18 and (ii) oxidation within the anode compartment of the
PEMFC stack itself.19 Both methods involve two reaction steps: absorption of CO on
the catalyst followed by oxidation of the absorbed CO to CO2, which is desorbed. In
method (i) the catalyst is the anode of an electrochemical cell in which surface oxygen
is generated by passage of an electric current. In method (ii) the first step is absorption
of CO directly on the supported Pt of the fuel‐cell anode, and the second step is
instigated by momentarily disconnecting the load on the fuel cell and applying a
positive potential to the anode. The latter serves to generate oxygen directly on the
surface of the anode electrocatalyst (essentially by electrolysis of water within the cell,
i.e., the cell is switched from fuel cell to electrolyser mode). Atomic oxygen on the anode
directly oxidizes the CO to CO2, and when this process is complete, the fuel‐cell mode
is resumed. Electrochemical oxidation within the cell is a simple concept but has its
own challenges, namely, increased catalyst degradation caused by switching between
oxidizing and reducing conditions and reduced cell efficiency since the oxidation to
CO2 effectively imposes an additional parasitic power load on the fuel‐cell stack.
Pressure swing adsorption (PSA) is a further method of hydrogen purification that can
be applied to reformates. In this process, the reformer product gas is passed into a
reactor that contains a material that preferentially absorbs hydrogen. After a set time, the
reactor is isolated and the feed gas is diverted into a parallel reactor. The first reactor is
then depressurized and thereby enables pure H2 to desorb from the material. The process
is repeated and the two reactors are alternately pressurized and depressurized.
17 Balasubramanian, S, 2011, Electrochemical oxidation of carbon monoxide in reformate hydrogen for
PEM fuel cells, PhD Thesis, University of South Carolina.
18 US Patent 6245214 B1, Electro‐catalytic oxidation (ECO) device to remove CO from reformate for fuel cell
application.
19 US Patent 5601936, A Method of operating a fuel cell.
Fuels for Fuel Cells
10.5
Membrane Developments for Gas Separation
As an alternative to PSA and the other methods of gas purification described in
Section 10.4.11, research is in progress to find an effective means to separate hydrogen
from CO2 subsequent to the WGS reaction. In particular, it is desirable to separate the gases
while hot and so conserve heat energy. The work is directed primarily towards the use of
membranes that are selective to the diffusion of H2 (a small molecule) while excluding
CO2 and other species. By using ceramic membranes, it should be possible to effect the
separation at close to the temperature of the WGS reaction or even the reforming reaction.
Various membrane reformers and membrane WGS reactors have been proposed, and
although these are already applied to some extent, there remains great scope for improving
the performance and lowering the costs of all types of membrane used for gas separation.
In general, membranes may be classified as (i) non‐porous, e.g., membranes based on
metals, alloys, metal oxides or metal–ceramic composites, or (ii) ordered microporous
materials, e.g., dense silica, zeolites and polymers.
10.5.1
Non‐Porous Metal Membranes
Metal‐based, non‐porous membranes can produce an H2 stream of very high purity that
can be used directly in a fuel cell. The separation process relies on the ability of the metal
to allow only the diffusion of H2. The permeation of H2 through metals such as palladium
and its alloys is thought to proceed via several steps, namely, adsorption of molecular
H2, dissociation to the monatomic form, ionization, diffusion of the hydrogen ions or
atoms under a concentration gradient through interstices within the metal lattice,
reassociation and, finally, desorption. It is the surface properties of palladium and its
alloys that give rise to high catalytic activity for absorption, dissociation and desorption.
Hydrogen flux density (and hence membrane performance) is a function of the inherent
diffusion characteristics of the material, and membranes above about 10 µm in thickness
are limited in performance by diffusion. In recent years, therefore, research has been
directed towards making thinner palladium and palladium alloy membranes. This also
has a benefit in terms of cost reduction. Very thin films (<10 µm) of palladium or its
alloys can be supported on a porous metal or ceramic substrate. Improved performance
can be obtained by raising the metal temperature. Indeed, it has been found advantageous
to maintain palladium and its alloys above the critical temperature (293°C for pure
palladium) to avoid stresses otherwise caused by the interaction between the two forms
of palladium hydride (PdH) with different crystal structures that can coexist at low
temperatures. Such interactions can lead to mechanical degradation and failure of the
metal — a process known as ‘hydrogen embrittlement’.
Unlike palladium and its alloys that are crystalline in nature, a class of amorphous alloy
membranes is emerging for H2 separation at high temperature. These are composed
primarily of nickel and early transition metals (i.e., titanium, zirconium, niobium, hafnium
and tantalum). Such alloys have a random atomic configuration, and, as with crystalline
alloy membranes, H2 migrates through the alloy via interstices between the metal atoms.
The first nickel‐based amorphous alloy membrane to be reported was Ni64Zr36. The
permeability of this alloy is about 10% that of palladium under similar conditions, and its
operation is limited to relatively low temperatures, typically below 400°C.
293
294
Fuel Cell Systems Explained
10.5.2
Non‐Porous Ceramic Membranes
A further class of non‐porous membrane is based on proton‐conducting metal oxides
from the perovskite family. These ceramic materials have the general formula ABO3
or A1 x A x B1 y B y O3 where x and y are fractions of dopants in the A‐ and B‐sites,
respectively, and δ is the number of oxygen vacancies. Considerable research has
been undertaken on SrCeO3 and BaCeO3 that have been variously doped with
trivalent cations such as those of yttrium, ytterbium or gadolinium. These oxides can
operate at much higher temperatures (up to 800°C) than metal membranes, which
makes them suitable for use in membrane reformer reactors. Unfortunately, however,
the oxides are difficult to make and suffer from low mechanical strength, poor H2 flux
and low chemical stability in the presence of both CO2 and water. These features
make them less applicable for medium‐temperature separation of H2, for example, in
conjunction with a WGS reactor. By contrast, zirconates (e.g., barium zirconate
(BaZrO3)) have better chemical stability but lower protonic conductivity. Cerate–
zirconate solid solutions that combine the favourable attributes of each have
been developed.
Ceramic–metallic (‘cermet’) materials have also been investigated as H2 separation
membranes. As with the MIEC oxides discussed in Section 9.2.3.3, Chapter 9, the cermet
can exhibit improved performance through the incorporation of an electronically
conducting metal to enhance the proton conductivity of the ceramic oxide.
10.5.3
Porous Membranes
A porous hydrogen separation membrane usually consists of a thin layer of a microporous
sieve material such as silica, carbon or zeolite on a thicker and highly porous support.
To maximize the flux, the microporous material is made much thinner than the
dense membranes discussed earlier, i.e., typically of the order of tens to hundreds
of nanometres. Hydrogen is transported predominantly through the pores of the
membrane by molecular diffusion, which is a purely physical process with a performance
that is determined by the pore diameter of the membrane. To separate H2 effectively,
the pores must be less than 1 nm in diameter.
A variety of established manufacturing techniques can be used to fabricate microporous
membranes on either metallic or ceramic macroporous components. Silica membranes,
for example, are made by coating the surface of a porous material with a silica‐based
chemical precursor. Two procedures can be employed. One method involves the dipping
of a suitable porous support into a sol–gel that contains a silica precursor such as
tetramethyl orthosilicate or tetraethyl orthosilicate. Multiple applications are usually
required to eliminate pinhole defects in the membrane. Similar precursors are used in
the alternative method but are applied by means of chemical vapour deposition (CVD).
In either procedure, the silica that is formed is partially densified by heating to form a
xerogel, with the desired pore‐size distribution. If heated further, the micropores will
close and render the membrane ineffective. For this reason, microporous membranes
are usually limited to gas separations below about 600°C.
Membrane separators are usually produced in ‘tube‐and‐shell’ configurations that are
assembled in multi‐tube modules for efficient distribution of feed and product gases;
see Figure 10.7.
Fuels for Fuel Cells
Hydrogen-selective
membrane film
CO
CO
H
Hydrogen
Carbon dioxide
H
H
H
H
H
H
20 nm
Hydrogen
Carbon dioxide
Figure 10.7 Schematic of a membrane reactor.
10.5.4
Oxygen Separation
The membranes so far discussed are used to separate H2 from a gas mixture. Also
relevant for fuel‐cell systems are membranes that are able to separate oxygen from air.
If oxygen, rather than air, is supplied to a CPO reactor or autothermal reformer, the
syngas product contains no nitrogen. This is beneficial on two counts. First, the absence
of nitrogen ensures that no ammonia can be formed in the processing steps. Ammonia,
which is discussed further in Section 11.6.3, Chapter 11, is harmful to PEMFCs and can
cause permanent degradation. Nitrogen by itself is generally a diluent in the fuel gas
stream and passes through most stacks without undergoing any chemical reaction.
Nevertheless, the diluting effect of nitrogen on the reactive gases inevitably reduces the
performance of all fuel cells, as indicated by the Nernst equation. Nitrogen also lowers
the effectiveness of other processes that may utilize the syngas, such as GTL.
Industrially, oxygen is separated from air by cryogenic processes or by PSA in which
the gas is absorbed on beds of zeolite that are alternatively pressurized or depressurized. Pressure swing absorption is also used for oxygen concentrators used in medicine.
Oxygen‐ion‐conducting membranes, as used in SOFC electrolytes, can also be employed
for oxygen separation and are starting to be adopted in fuel‐cell systems. To avoid a
charge distribution developing across the membrane caused by the migration of ions,
an MIEC material is preferred for such duties.
10.6
Practical Fuel Processing: Stationary Applications
10.6.1
Industrial Steam Reforming
Before practical fuel processing for fuel‐cell systems is considered, it is instructive to
consider the operation of a typical industrial steam reforming plant. Such facilities have
been built for many decades to provide H2 for both oil refineries and chemical plants
295
296
Fuel Cell Systems Explained
(mainly to produce ammonia for the fertilizer industry). Industrial reformer systems
usually produce between 7 and 30 million normal cubic metres (Nm3) of H2 per day.
The systems, which consist of a number of tubular reactors packed with catalyst pellets,
operate at temperatures up to 850°C and pressures up to 2500 kPa. The reactors are
generally around 12 m in length and must be made from expensive alloy steels to endure
both the high temperatures and the reducing gas conditions. Such reformers can be
scaled down reasonably easily to give H2 outputs of some 0.1–0.3 million Nm3 day−1.
It was noted in Section 2.3, Chapter 2, that the LHV for the enthalpy of combustion
of H2 is −241.8 kJ mol−1. Hydrogen supplied at a rate of 1.0 Nm3 h−1 when combusted
will therefore produce about 3 kW of heat energy. If the H2, rather than being combusted, is fed to a fuel‐cell system that has an overall electrical efficiency of 40% (LHV),
then clearly the energy produced by such a fuel‐cell systems fed with 1.0 Nm3 h−1 of H2,
for example, would be 0.4 × 3 = 2.4 kW.
Unfortunately, scaling down an industrial reformer that is sized for several million
Nm3 per day to one that provides only a few Nm3 h−1 of H2 is not a practical proposition.
Conventional tubular reformers are expensive because of the need to run at high
temperatures and pressures, and they are large in terms of footprint area and weight.
Accordingly, alternatives have to be provided to suit the much smaller demands of fuel‐
cell systems.
10.6.2
Fuel‐Cell Plants Operating with Steam Reforming of Natural Gas
For stationary power plants that employ PEMFC and PAFC stacks, steam reforming of
natural gas is the preferred option for generating H2 because it gives high fuel‐conversion
efficiency for the system as a whole. The technology has been applied for many years in
facilities of between about 50 kW and several MW. In both PEMFC and PAFC systems,
the sulfur removal can be achieved by HDS. For PAFC systems, two stages of shift
are required for lowering the level of CO in the reformed gas. In the case of a PEMFC,
however, a further CO removal step will also be necessary.
With such systems, a degree of process integration is required so that heat from the fuel
cell is utilized for various preheating duties. The chemical processes (desulfurization,
steam reforming, WGS and CO removal) each take place at a different temperature.
Consequently, there are a number of temperature changes to be made. The minimum
requirements are as follows:
●
●
●
●
●
●
Initial heating of the dry fuel gas to approximately 300°C prior to HDS.
Further heating of gas and steam prior to steam reforming at 600°C or higher.
Cooling of the reformer product gas to approximately 400°C for the high‐temperature
WGS reaction.
Further cooling to approximately 200°C for the low‐temperature WGS reaction.
Temperature adjustment prior to entry into the CO removal step or directly into the
fuel cell (depending on the fuel‐cell type).
Heating of water to produce the steam required for the steam reformer.
In addition to these six temperature changes, steam reforming demands high‐
temperature heat. This requirement can be met by burning the anode exhaust gas,
which always contains some unconverted fuel. Further heating of the anode exhaust gas
may also prove advantageous. Similarly, preheating the air to the burner will yield a
higher combustion temperature. In such fuel‐cell systems, therefore, some gases have to
Fuels for Fuel Cells
Steam in
Exhaust
gas out
120
400
To fuel cell
Anode
A
Air in
300
20
B
220
250
Steam reformer
150
850
Lowtemperature
shift
converter
Burner
400
In the fuel
cell most,
but not
all, of the
hydrogen
is used
250
800
C
D
650
220
Steam
450
Desulfurizer
Hightemperature
shift
converter
250
280
450
650
E
20
Natural
gas
F
380
220 Anode
exhaust
gas
Figure 10.8 Diagram of a fuel‐processing system for a phosphoric acid fuel cell. The numbers indicate
approximate temperatures (°C).
be heated and others have to be cooled. Heating and cooling can be combined using
heat-exchangers (see Section 7.2.3, Chapter 7).
A simplified flow diagram of a fuel‐processing system for a PAFC powered by natural
gas is shown in Figure 10.8. The PAFC requires reformed fuel gas at about 220°C, with
a CO content below about 0.5 vol.%. The following is an explanation of the process:
●
The natural gas enters the fuel processor at around 20°C and is heated in heat‐
exchanger E to a temperature that is suitable for desulfurization (280°C). Steam,
sufficient for both the reforming and the WGS reaction, is then mixed with the
desulfurized fuel. The steam–methane mixture is further heated by heat‐exchanger
297
298
Fuel Cell Systems Explained
●
●
●
●
●
●
C before being fed to the steam reformer. Here, it is heated to around 850°C by the
burner and is converted to a syngas product. Note that the syngas also contains
some unreacted steam.
The syngas then passes through the other side of heat‐exchanger C and loses heat to
the incoming fuel gas. Further heat is lost to both the incoming gas at E and the anode
exhaust gas at F.
The gas is now sufficiently cool for the first WGS converter, where the majority of the
CO is converted to CO2. At D, the gas is further cooled by giving up its heat to the
incoming steam and then passes to the low‐temperature WGS converter for conversion
of the remaining CO to CO2. The final cooling is accomplished at B, where the incoming
steam is heated.
The H2‐rich fuel gas is then sent to the PAFC stack. Here most, but not all, of the H2
is converted to electrical energy. The anode exhaust gas, still at about 220°C, is sent to
heat‐exchanger F where it is preheated prior to reaching the burner.
The burner is also fed with air, which will have been preheated by heat‐exchanger
A through the use of energy from the burner exhaust gas.
The steam arriving at about 120°C at heat‐exchanger B can be generated from water
by using heat from the cooling system of the fuel cell.
The burner exhaust gas, still very hot, can also be employed to raise steam to power
any compressors that are required to drive the process.
There are many other possible ways of configuring the gas flows and heat-exchangers to
achieve the desired result, but the process flow diagrams of commercial systems are
usually proprietary. A stationary fuel‐cell system is analysed further in Section 12.4.4,
Chapter 12.
10.6.3
10.6.3.1
Reformer and Partial Oxidation Designs
Conventional Packed‐Bed Catalytic Reactors
Early PAFC plants, such as those developed by United Technologies Corporation
(UTC), International Fuel Cells and Fuji Electric in the 1980s, employed fairly
traditional designs of fuel processor that consisted of fixed catalytic beds for the
desulfurization, steam reforming and WGS reactions. Heat from burning natural
gas in a conventional burner is transferred mainly by radiation to the reformer
reactor(s) that, as in a typical large‐scale refinery installation, is(are) operated at
above 850 C.
In 1989, researchers at WS Reformer GmbH discovered that stable combustion of
natural gas with air could be obtained by increasing the throughput of the burner
and recirculating the resulting product gas. The procedure became known as
flameless oxidation or FLOX™ and, compared with conventional burner designs,
offered the advantages of lower and more uniform temperatures throughout the
combustor and a significantly lower level of nitrogen oxides (NO x) in the burner
off‐gas. The FLOX™ concept was applied by WS Reformer GmbH to systems for
generating H2 for the fuel‐cell buses that were undergoing trials in the HyFLEET
Clean Urban Transport Europe (CUTE) programme. FLOX™ reformers have
also been demonstrated with the 1‐kW high‐temperature PEMFC manufactured
by Serenegy in Sweden.
Fuels for Fuel Cells
10.6.3.2
Compact Reformers
The desulfurizer, WGS reactors and CO clean‐up systems in all of the reformer systems
may be packed‐bed catalytic units of traditional design. In many cases, the pellets or
extrudates that are the common forms of catalyst used in the petrochemical industry
have been replaced by coated ceramic monoliths for fuel cells. With respect to the
design of the reformer reactor for fuel‐cell systems, several novel features are being
pursued, particularly in integrating some of the heat‐transfer duties.
A compact reformer produced by Haldor Topsøe for PAFC systems is shown in
Figure 10.9. In this particular design, heat for the reforming reaction is provided by
combustion of the lean anode exhaust gas, which may be supplemented with fresh fuel
gas. Fuel is combusted at a pressure of some 450 kPa in a central burner that is located in
the bottom of a pressure vessel. Feed gas is passed downwards through the first catalyst
bed where it is heated to around 675°C by convection from a combined countercurrent
of the combustion products and the reformed product gas. On leaving the first bed of
catalyst, the partially reformed gas is transferred through a set of tubes to the top of the
second reforming stage. The gas flows down through the catalyst and is heated typically
to 830°C by convection from the co‐currently flowing combustion products and also by
radiation from the combustion tube. The combination of co‐current and countercurrent
heat transfer helps moderate the temperature of the reactor, an important consideration
in high‐temperature reformer design. The advantages of such a reformer for fuel‐cell
applications are (i) small size and suitability for small‐scale use, (ii) pressurized
combustion of lean anode exhaust gas gives good process integration with the fuel cell,
Reformed
fuel out
Fuel to be
reformed
Flue gas
out
First catalyst
bed
Second catalyst
bed
Anode off-gas
containing unreacted
fuel
Burner
Air
Figure 10.9 Haldor Topsøe heat‐exchange reformer.
299
300
Fuel Cell Systems Explained
(iii) improved load following and (iv) lower cost. In addition to Haldor Topsøe, several
companies have been developing reformers of this type, namely, International Fuel Cells
(now Doosan Fuel Cell), Ballard Power Systems, Sanyo Electric, Osaka Gas and
ChevronTexaco.
10.6.3.3
Plate Reformers and Microchannel Reformers
A plate reformer is built up of a number of box-like reactors stacked one on top of the
other. Thin metal plates separate each reactor compartment. The compartments are
alternately filled with suitable catalysts to promote the combustion and steam reforming
reactions. In another approach, each separating plate is coated with a steam reforming
catalyst on one side and a combustion catalyst on the other. The heat from the combustion
reaction is used to drive the reforming reaction. Plate reformers have the dual advantage
of being very compact and offering a means of maximizing heat transfer. The use of a
combustion catalyst enables gases with low heating values (e.g., anode exhaust gases)
to be burnt without the need for a supplementary fuel. Plate reformers were first
developed by Ishikawajima‐Harima Heavy Industries Co., Ltd (IHI) in the early 1980s;
the catalyst was in the form of spherical pellets located on either side of the heat‐
exchanger surface.20 Gastec, Plug Power, Osaka Gas and several other companies have
since built plate reformers for fuel‐cell systems.
The most advanced types of plate reformer employ compact heat‐exchanger hardware
on which the catalyst is coated directly in the form of a thin film with a thickness of
a few microns.21 The concept is shown in Figure 10.10. Such devices were developed
in the United States by researchers at Pacific Northwest National Laboratory who have
demonstrated a 1‐kW steam reformer and at Argonne National Laboratory who
are developing a reformer with a monolithic catalyst for the processing of diesel. Plate
reformers for methanol have been constructed by several organizations that include
CH4 + O2
CO2 +H4O
Active sites
Porous
catalyst
support
10 – 50 μm
1 mm
Heat
10 – 50 μm
Stainless-steel
substrate
Porous
catalyst
support
CH4 + H2O
H2 + COx
Figure 10.10 Plate or microchannel reformer concept. Catalyst is coated as thin film onto one or both
sides of a heat‐exchange material.
20 Hamada, K, Mizusawa, M and Koga K, 1997, Plate reformer, US Patent No. 5,609,834.
21 Goulding, PS, Judd, RW and Dicks, AL, 2001, Compact reactor, Patent No. WO/2001/010773.
Fuels for Fuel Cells
(a)
(b)
(c)
Figure 10.11 Experimental compact reformer reactors. (a) and (b) Pacific Northwest National Laboratory
liquid fuel reformer (courtesy of Pacific Northwest National Laboratory). (c) Proof‐of‐concept Advantica
natural gas reformer, a diffusion‐bonded multichannel reactor block (3 × 3 × 10 cm). (Source: Courtesy of
Advantica Technologies Ltd.)
IdaTech, Mitsubishi Electric, InnovaTek Inc., NTT Telecommunications Laboratory
and Honeywell. Examples of the hardware are shown in Figure 10.11.
A microchannel reactor (MCR) is another term for the compact reactor technology
that could be applied to other units of a fuel processor such as the fuel vapourizers and
the gas clean‐up reactors. Unfortunately, however, MCR systems suffer from two
notable drawbacks, namely, (i) plugging of the channels due to catalyst degradation
and carbon deposition, and (ii) the catalyst is incorporated into the reactor for life and
therefore cannot easily be replaced when it becomes degraded.
10.6.3.4
Membrane Reactors
An example of a membrane reactor developed by a commercial company is the ion transport
membrane (ITM) technology developed by Air Products that combines air separation and
high‐temperature generation of synthesis gas (via autothermal reforming) in a single
ceramic membrane reactor. The ITM Syngas process employs a planar membrane of
MIEC oxides. In operation, oxygen from a hot air stream is reduced at one surface of the
ITM membrane to oxygen ions, which diffuse through the membrane under a chemical
301
302
Fuel Cell Systems Explained
Scanning electron
microscope
cross-section
H2 separation membrane
Diffusion barrier (YSZ)
H2 separation
membrane
(Pd/Al2O3)
H2
Catalyst support
Ni/ YSZ
H2
H2
Reactor body
Catalytic support
20 μm
H2
H2
H2
H2O + CH4
CO + CO2 + H2O
H2
H2
H2
H2
H2
H2
Figure 10.12 Concept of membrane reformer system under development by Tokyo Gas.
potential gradient. At the opposite surface of the membrane, the oxygen partially oxidizes
a preformed mixture of hot natural gas and steam to form syngas.
Praxair is exploring a similar strategy that differs mainly in that a tubular membrane
is used. Although such technologies have been targeted for GTL operations, there is no
reason in principle why they should not be applied to fuel‐cell systems.
In 2009, Tokyo Gas demonstrated a membrane reformer system that was capable
of producing H2 from natural gas at a rate of 40 Nm3 h−1. The reactor system,
including insulation, measured 1200 × 50 × 1350 mm and was comprised of 112
tubular membrane modules, each made of porous stainless steel onto which was
coated a 20‐µm layer of palladium. The H2 production efficiency was claimed to be
over 81%, which is significantly higher than that obtained with a conventional
reformer reactor. Current research by Tokyo Gas is focused on using a thin‐film
palladium–silver alloy that is deposited on yttria‐stabilized zirconia that acts as
both a diffusion barrier for carbon oxides and a support to the reformer catalyst.
The work programme to 2020 is aimed mainly at H2 production systems for
FCVs and is based on the use of a tubular reactor, as illustrated schematically in
Figure 10.12.
10.6.3.5
Non‐Catalytic Partial Oxidation Reactors
Non‐catalytic partial oxidation (NCPO) is applied industrially by Texaco and Shell for
the conversion of heavy oils to syngas. In the Shell process, liquid fuel is fed to a reactor
together with oxygen and steam. A partial combustion takes place in the reactor and
yields a product at around 1150°C. It is this high temperature that poses a particular
problem for conventional POX. The reactor has to be made of expensive materials, and
the product gas needs to be cooled to allow unreacted carbon material to be separated
from the gas stream. The high temperature also means that expensive materials of
construction are required for the heat-exchangers. In addition, the effluent from the
reactors invariably contains contaminants (e.g., sulfur compounds) as well as carbon
Fuels for Fuel Cells
and ash, all of which require proper disposal. Therefore, given its extra cost and complex
operation, NCPO has not been a preferred option for fuel‐cell applications.
10.6.3.6
Catalytic Partial Oxidation Reactors
Catalytic partial oxidation reactors can be very simple in design given the requirement
for only one bed of catalyst into which the fuel and oxidant (usually air) are injected.
Often steam is added as well, in which case some conventional reforming also occurs.
As mentioned in Section 10.4.7, the combination of CPO and steam reforming is usually
referred to as ‘autothermal reforming’ because there is no net heat supplied to, or
extracted from, the reactor. All of the heat for reforming is provided by partial
combustion of the fuel. Depending on the nature of the fuel and the application, two
types of catalyst are sometimes used — one primarily for the CPO reaction and the
other to promote steam reforming.
The HotSpot™ reactor, developed during the late 1990s, is an example of a CPO
reactor. The technology was promoted by Johnson Matthey as a means of generating H2
from natural gas for small‐scale stationary systems, as well as for the reforming of liquid
fuels on vehicles. The reactor employed a platinum–chromium oxide catalyst on a
ceramic support. Three reactors are shown in Figure 10.13. As its name implies, the
novel feature of the reactor was the hotspot caused by point injection of the air–
hydrocarbon mixture through a narrow tube inserted into the centre of the catalyst bed.
The arrangement eliminated the need for preheating the fuel gas and air during operation, although, for start‐up on natural gas, the fuel had to be preheated to around 500°C.
Alternatively, the reactor could be started from ambient temperature by introducing an
initiating fuel such as methanol or an H2‐rich gas. These fuels are oxidized by air at
ambient temperature over the catalysts and thereby serve to raise the bed to the
temperature required for natural gas to react (typically over 450°C).
Figure 10.13 Johnson Matthey HotSpot™ reactor. These were made in different forms for methanol,
methane or gasoline processing. (Source: By courtesy of Johnson Matthey plc.)
303
304
Fuel Cell Systems Explained
10.7
Practical Fuel Processing: Mobile Applications
The motivation for the reforming of fuel on vehicles was probably the world oil crisis of
1974 that stimulated support for the advancement of all types of fuel cell, as well as
interest in the possibility of a ‘hydrogen economy’. The rapid development of the PEMFC
during the 1990s and its uptake by leading vehicle manufacturers led to some impressive
research and development programmes being mounted by DaimlerChrysler, General
Motors, Ford and others in collaboration with manufacturers of PEMFC stacks.
Methanol was proposed as a suitable energy carrier because, in addition to being a liquid
and therefore readily transportable, it reacts with steam over a catalyst at relatively low
temperatures. On‐board reforming of methanol was therefore pursued during the 1990s
by several groups and organizations that included Johnson Matthey, DaimlerChrysler,
General Motors, Ballard Power System, Nissan and Toyota. Many demonstration
vehicles were built, and towards the end of the decade, Arthur D. Little, ExxonMobil,
Nuvera and Shell had also investigated catalysts and processes for on‐board reforming
of gasoline.
As the 21st century arrived, many questioned the wisdom of on‐board fuel reforming.
This was due not only to the technical difficulties associated with the time and energy
required at start‐up and the poor transient response but also to the high costs involved.
In addition, various ‘well‐to‐wheel’ studies carried out in Europe and North America
showed that on‐board gasoline reforming for an FCV did not achieve a particularly
high overall energy‐conversion efficiency in comparison with an internal combustion
engine vehicle (ICEV) or a hybrid electric vehicle (HEV). In 2004, the US DOE brought
the various stakeholders together for a go/no‐go decision for on‐board reforming.22
The DOE committee unanimously decided to abandon all support for on‐board reforming projects, an outcome that was generally accepted worldwide.
The only notable exceptions to the DOE decision are the on‐board fuel processors
that form a component of auxiliary power units (APUs) for heavy‐duty trucks, military
vehicles and recreational vehicles such as campervans. These all have significant power
requirements even if the vehicle is not moving, and they make a sizeable contribution to
global emissions and wasted energy. For example, idling diesel vehicles (including
trailers and buses) are estimated to burn a billion gallons of diesel fuel every year in the
United States alone. Diesel fuel or heavier logistic fuel can be converted to a syngas in a
CPO reactor, and, apart from sulfur removal, this facility alone is sufficient to fuel an
on‐board system for generating power with an SOFC. Any other fuel cell would be too
problematic in terms of fuel processing for this application. A combination of CPO and
SOFC is therefore an attractive proposition for producing the auxiliary power for
heavy‐duty trucks and similar sizes of vehicle. Especially when combined with some
battery storage, a CPO–SOFC need not exhibit fast dynamic response as the auxiliary
systems demand a fairly constant power load.
Delphi Corporation has been one of the leading proponents of on‐board SOFC–APU
technology with an integrated fuel processor. The company worked in partnership with
22 DOE Team Decision Report, August 2004, On‐board fuel processing go/no‐go decision.
Available online: http://www1.eere.energy.gov/hydrogenandfuelcells/pdfs/committee_report.pdf
(accessed on 27 September 2017).
Fuels for Fuel Cells
BMW, Los Alamos National Laboratory, Battelle and Global Thermoelectric to develop
a diesel‐fuelled APU in the 1990s. In 2001, a gasoline‐fuelled APU was demonstrated by
BMW in a 7‐series sedan.
In addition to APU systems for vehicles, the processing of gasoline and diesel has
been suggested as a means of providing power for the hotel load on ships, i.e., the power
that continues to be required by marine craft when they are anchored in harbour.
Consequently, a large body of literature exists on the steam reforming and CPO of these
fuels.23 The catalysts are much more demanding than those for processing simpler fuels
such as natural gas or methanol. Higher temperatures are necessary, and a greater
resistance to fouling by carbon or sulfur generally rules out simple supported nickel
catalysts. In general, the steam reforming of fuels such as kerosenes involves the removal
of sulfur compounds and the use of high steam‐to‐carbon ratios. In the case of CPO, the
absence of steam gives rise to lower concentrations of H2 in the reformate and a higher
risk of carbon formation. The influence of both the oxygen‐to‐carbon ratio and the
highly exothermic CPO reaction has led to the development of catalysts that are more
tolerant of carbon formation. For these catalysts, the preferred active metals are
rhodium and ruthenium, supported on ceria. Compared with the more conventional
supported nickel catalysts, the platinum‐group metal is less active for the reactions
that form carbon, and the ceria provides an oxygen‐rich surface functionality for the
cracking of highly aromatic hydrocarbons.
10.8
Electrolysers
10.8.1
Operation of Electrolysers
Electrolysers use electricity to split water into hydrogen and oxygen. They are thus
the opposite of a fuel cell. The basic theory and the reactions taking place at the
electrodes are the same for electrolysers as for fuel cells — except that the reactions
are reversed. Different electrolytes can be used, just as for fuel cells, and in order
to minimize electricity consumption, it is important to choose an electrolyte of
maximum conductivity. Electrolyser technology was developed in the 1800s, and by
the beginning of the 20th century, there were more than 400 industrial water
electrolysis units in operation. In 1939, the first large water electrolysis plant with an
H2 output of up to 10 000 Nm3 h−1, built by the Norwegian company Norsk Hydro
Electrolyzers, went into operation.
Most industrial electrolysers employ an alkaline electrolyte, and, as with the AFC, this
is usually an aqueous solution of potassium hydroxide (30–40 wt.%). The solution must
be prepared from very pure water; otherwise impurities will accumulate during
electrolysis. The chloride ion, which is usually present in water, is particularly harmful
in that it causes pitting of the protective films that form on metal surfaces in alkaline
solutions.
Industrial alkaline electrolysers have been employed to generate hydrogen for diverse
applications that range from the hydrogenation of fats in the food industry to the
23 Schwank, JW and Tadd, AR, 2010, Catalytic reforming of liquid hydrocarbons for on‐board solid oxide
fuel cell auxiliary power, Catalysis, vol. 22, pp. 56–93.
305
306
Fuel Cell Systems Explained
cooling of large gas turbines engaged in central electricity generation.24 In principle,
electrolysers are well suited for use with electricity generated from renewable energy
sources such as wind, solar and hydropower. In practice, however, alkaline electrolysers
are designed to operate with a fairly constant power supply, and therefore the intermittent
nature of renewable sources may necessitate the development of specialized power
control and conditioning equipment. Commercial electrolysers are also not particularly
efficient. A large industrial plant capable of producing 500 Nm3 h−1 of H2 would demand
about 2.3 MW of power. With a capital cost of US$600–700 per kW, this makes the
generation of H2 by electrolysis uneconomic unless low‐cost electricity is available.
Using renewable electricity, it is estimated that H2 can be produced at around US$7–10
per kg, i.e., three to five times above that from fossil fuels.25
Traditional alkaline electrolysers may be constructed with either monopolar or bipolar
configurations. The monopolar (or ‘tank‐type’) unit consists of alternating positive and
negative electrodes that are held apart by microporous separators. The positives are all
connected together in parallel, as are the negative electrodes, and the whole assembly is
immersed in a single electrolyte bath or tank to form a unit cell. By contrast, a bipolar
unit uses a metal bipolar plate to join adjacent cells, as in a PEMFC stack. In a bipolar
alkaline electrolyser, the electrocatalyst for the negative electrode is coated on one face
of the bipolar plate, and that for the positive electrode of the adjacent cell is on the
reverse face. A series‐connected stack of bipolar cells forms a module that operates at a
higher voltage and lower current density than the monopolar design.
In an alkaline electrolyser, the reactions occur through the ionization of water into
protons and hydroxyl ions, i.e.,
H2O
H( aq )
OH( aq )
(10.14)
The OH( aq ) ions migrate to the positive electrode to produce electrons and release
oxygen:
4OH( aq )
2H 2 O O 2 4 e
(10.15)
Concomitantly, hydrogen is produced at the negative electrode:
2H( aq ) 2e
H2
(10.16)
The reactions are, of course, the opposite of those occurring in the AFC.
Despite widespread use of the alkaline electrolyser, much interest has been shown in
two other technologies that offer certain advantages, namely, the proton‐exchange
membrane (PEM) electrolyser and the high‐temperature steam electrolyser.
24 In a synchronous generator or alternator, hydrogen gas is circulated by blowers and fans through the
rotor and stator and then passed over cooling coils inside the generator casing. The coils carry oil or water
to extract heat from the circulating hydrogen. Hydrogen has the advantage of low density (7% that of air)
and high thermal conductivity (6.7 times that of air). It offers several advantages over air for cooling
alternators or generators, for example, it allows a machine of the same dimension to have 20–25% greater
output capacity, and hydrogen cooled alternators required 20% less active material (steel and copper) than
air-cooled machines.
25 The Hydrogen Economy: Opportunities, Barriers and R&D Needs, 2004, National Academic Press,
Washington, DC. ISBN: 978‐0‐309‐09163‐3.
Fuels for Fuel Cells
The first PEM electrolyser was produced by General Electric in 1966. The basic structure is the same as the PEMFC, although the electrodes have different requirements. The
reactions for water electrolysis in an acid electrolyte, such as a PEM, are as follows:
At the negative electrode:
4H
4e
2H2
(10.17)
At the positive electrode:
2H2O
O2 4H+ 2e
(10.18)
These reactions are the opposite of those shown in Figure 1.3.
One reason for the success of PEM electrolysers is that many of the problems associated with PEMFCs do not apply. Cooling is fairly trivial, as the water supplied to the
cathode can be pumped around the cell to remove heat. Water management, another
key problem of the PEMFC, is also massively simplified, as the positive electrode must
be flooded with water. Electrolysers using PEM membranes also offer advantages over
their alkaline counterparts in terms of:
●
●
●
●
●
Wide operating range capable of meeting large variations in demand.
Greater safety through the absence of alkaline solutions.
More compact design due to higher current densities.
Ability to operate at higher pressures.
Minimal maintenance required.
The H2 produced by such electrolysers will have a high purity but may have a high
humidity due to protons dragging water molecules through the electrolyte. Indeed, the
water content of H2 can be so high that condensation occurs and can prove to be a
serious issue if it is intended to store the H2 at pressure or as a solid‐state hydride.
Since the energy to compress water is less than that required to compress H2, the
high‐pressure PEM electrolyser (HPE) has been the subject of increasing development
in recent years. Hydrogen pressures of 12–20 MPa can be achieved and thus eliminate
the need for a gas compressor that can be both expensive to maintain and inefficient.
Stacks employed in HPEs are shown in Figure 10.14.
Theory predicts that elevating the temperature of electrolysis would improve
efficiency, since some of the energy used to split the water is provided as heat, and
would also reduce the overpotentials at the electrodes. Increasing the temperature substantially to 700–1000°C is possible if a sold ceramic electrolyte is used, e.g., stabilized
zirconia, as in the SOFC. The concept of such a high‐temperature ‘steam electrolyser’
has been investigated for many years in parallel with the development of the SOFC and
other high‐temperature electrochemical reactors.26
10.8.2
Applications
At first, it would seem to be perverse to use electricity to make H2 for use in fuel cells to
turn it back into electricity. In each conversion step there are losses in efficiency, such
that the ‘round‐trip’ efficiency of converting electricity to H2 and then back again can
26 Zahid, M, Schefold, J and Brisse, A, 2010, High‐temperature water electrolysis using planar solid
oxide fuel cell technology: a review, in Stolten, D (ed.), Hydrogen and Fuel Cells, pp. 227–231,
Wiley‐VCH, Weinheim.
307
308
Fuel Cell Systems Explained
Figure 10.14 ITM Power HGas electrolyser stacks, each operating at 8 MPa pressure.
(Source: Reproduced with permission of ITM Power.)
be as low as about 40%. If the objective, however, is to avoid the wastage of surplus
electricity (v.i.) through using it to make hydrogen that can be stored to meet a future
energy demand, then it may be possible to make an economic case for electrolysis. In
general, the H2 cost is a function of both the cost of electricity and the capital cost of the
associated equipment, i.e., the electrolyser and the compressor. At the small scale, solar
or wind power can be used to generate H2 for use in FCVs. In this case, the cost of
electricity is governed principally by the capital cost of the photovoltaic (PV) panels and
associated control and inverter equipment. Several manufacturers are providing systems
for generating H2 for FCV filling stations and include ITM Power, Hydrogenics and
Proton OnSite. Two examples are shown in Figure 10.15.
On a larger scale, electrolysis can be used to generate H2 from the excess or off‐peak
power generated by wind or solar farms that can then be stored for future use. The concept
has become known as power‐to‐gas (often abbreviated to P2G). The H2 produced in P2G
systems may be converted further to substitute natural gas (SNG) via methanation of
CO2 or to syngas and liquid fuels by the Fischer–Tropsch (GTL) process (v.s.). On the
other hand, the H2 may simply be compressed and blended with natural gas in the gas
transmission or distribution network.27 Natural gas pipelines can provide a strategic
means of bulk energy storage and can accept low concentrations of hydrogen. In Germany,
for example, the capacity of the natural gas network is more than 200 000 GWh, which is
sufficient to satisfy the national energy demand for several months. By comparison, the
27 Whereas substitute natural gas can be blended directly into a gas transmission network, there is a
limiting concentration to which hydrogen can be blended, typically a few per cent by volume. The allowable
concentration of hydrogen varies according to the country and point of injection.
(a)
(b)
Figure 10.15 Hydrogen filling stations using (a) ITM Power PEM electrolysers coupled with wind
turbine at the Advanced Manufacturing Park, South Yorkshire, United Kingdom. (b) Hydrogenics
HySTAT alkaline electrolysers 130 kg day−1 dispensing 70 MPa H2, Stuttgart, Germany.
(Source: Reproduced with permission of ITM Power.)
310
Fuel Cell Systems Explained
capacity of all German pumped‐storage utilities amounts to only about 40 GWh. Note
that natural gas pipelines cannot be used safely to distribute pure hydrogen without being
modified to ensure against progressive degradation of the materials employed in
construction of these networks. High carbon steels employed in transmission systems can
be prone to hydrogen embrittlement and decarburization when subject to pressure cycling
in pure hydrogen, and polyethylene used in low‐pressure distribution systems is not
compatible with hydrogen.
For P2G systems, the PEM electrolyser is favoured because of its ability to operate at
sufficiently high pressure so that the hydrogen produced can be blended directly into the
natural gas network. A further advantage is the faster response of the PEM electrolyser
compared with an alkaline system. As more renewable energy sources are fed into a
traditional electricity supply network, there will come a point where the intermittent
renewable power is out of step with the demand from consumers. Traditional turbogenerators are unable to react sufficiently rapidly to balance the load on the network with
the consequence that voltages on the network may rise or fall and the AC frequency may
shift outside the prescribed limits, with potentially catastrophic results. It is generally
recognized that some form of grid storage will therefore be required as renewable power
becomes more widespread. Unfortunately, however, most storage options, such as
rechargeable batteries, are very expensive per kWh or, in the case of pumped hydro,
respond too slowly to changes in demand. To address the load‐balancing issue for power
networks, ITM Power has produced the HGas rapid‐response PEM electrolyser shown
in Figure 10.16. The HGas system is modular and can be supplied in a range of sizes
upwards from 70 kW (producing 20 kg H2 day−1). The HGas system can be turned on
within a few seconds to deliver H2 at 8 MPa that can be either stored or fed directly into
the natural gas network without further compression.
Over the past 20 or so years, many projects have been initiated to demonstrate the
feasibility of using PV or wind to supply an electrolyser to generate H2 that can then
be stored, either as compressed gas or in the form of a hydride. A fuel‐cell system —
typically an AFC, PEMFC or PAFC — is coupled with the hydride. In most cases, lead–
acid batteries have also been incorporated into the systems to assist with load levelling.
Example projects were the HARI demonstration undertaken by Beacon Energy and
Loughborough University in the United Kingdom and the RES2H2 project conducted
in Greece. Some common issues with these systems were identified:
●
●
●
●
●
●
Hydrogen from an alkaline electrolyser must be purified to be acceptable for hydride
storage (see Section 11.5, Chapter 11).
Steps need to be taken to minimize contamination of the H2 when the equipment is
in standby mode.
The electrolyser can cope with intermittent power from PV or wind turbines,
provided a battery is used as a storage buffer.
Careful process integration is required to minimize heat loss and ensure high overall
efficiency.
The requirement for water purification could be largely reduced, or even eliminated,
by recycling product water from a PEMFC.
Problems are more likely to arise with the auxiliary mechanical equipment (e.g., water
demineralizer, air compressor, inert gas supply) than with the electrolyser and storage
components.
Fuels for Fuel Cells
(a)
(b)
Figure 10.16 ITM Power HGas rapid‐response electrolyser: (a) installed in a P2G system in Frankfurt,
Germany and (b) stack of PEM cells. (Source: Reproduced with permission of ITM Power.)
311
312
Fuel Cell Systems Explained
10.8.3
Electrolyser Efficiency
The thermodynamically ‘reversible’ voltage Vr for water electrolysis under isothermal
conditions and at standard temperature (298.15 K) and pressure (101.325 kPa) is 1.229 V.
This value decreases almost linearly with increasing temperature to 1.0 V at 573 K
(300°C). The concomitant decrease in free energy, ΔG, is largely offset by an increase in
the entropy term, TΔS, so that the enthalpy of reaction, ΔH, is almost independent of
temperature. The efficiency of an electrolyser is calculated in almost the same way as
that for a fuel cell. If Vc is the operating voltage for one cell of a fuel‐cell stack, then it
was shown in Section 2.4, Chapter 2, that the efficiency (on a higher heating value
(HHV) basis) is given by:
Vc
1.48
(10.19)
In the case of an electrolyser, the formula is simply the inverse of this expression,
namely:
1.48
Vc
(10.20)
Apart from there being no issue of fuel crossover, the voltage losses in electrolysers
follow exactly the same pattern as those for fuel cells that are described in Section 3.3,
Chapter 3, i.e., activation losses (overpotentials) at the positive and negative electrodes
and the total resistive (‘ohmic’) losses within the electrodes and electrolyte. An alkaline
electrolyser when operating at 90°C and atmospheric pressure with non‐precious metal
electrodes typically requires 2.1 V overall to yield a current density of 200 mA cm2. Because
the problems of cooling and water management are so much more easily solved in the
PEM electrolyser, the performance of this technology is routinely higher than that for an
alkaline electrolyser; in fact, it can match that of the best PEMFCs, with current densities
of around 1.0 A cm−2 usually being achieved. To keep capital costs low, it is necessary to
operate electrolysers at as high a current density as possible, but, as with PEMFCs, such
practice has to be traded against lower cell efficiency. Industrial alkaline electrolysers are
generally 60–75% efficient, whereas best practice with small‐scale systems is claimed to
be closer to 80–85%. The German HOT ELLY high‐temperature electrolyser, which uses
a zirconia electrolyte, reaches an efficiency of over 90%. Despite this high efficiency, high‐
temperature electrolysers still produce H2 at about four times the cost of that obtained
from the steam reforming of natural gas.
10.8.4
Photoelectrochemical Cells
Sunlight may be harvested by PV cells to generate DC power for the electrolysis of
water. Photovoltaic solar cells, which are becoming widely used in their own right for
power generation, are formed from a thin layer of a semiconductor material, such as
silicon. The material is doped in such a way that one side is negatively charged (n‐type)
and the other positively charged (p‐type). When light strikes the n‐type semiconductor,
loosely held electrons are released, and, if a current-collector is attached, these electrons
can be sent via an external circuit to the p‐type side where they will be accepted by
Fuels for Fuel Cells
vacancies or ‘holes’ in the material. The resulting flow of current can be utilized, but the
voltage created by a single p–n junction is low, and because not all of the energy in the
light is captured (i.e., only a part of the spectrum has sufficient energy to release
electrons from the ‘conduction band’ to the ‘valence band’), the efficiency of PV cells
tends to be low. Other semiconductors that have been employed in PV cells include
gallium arsenide (GaAs), cadmium telluride (CdTe) and copper indium gallium
diselenide (Cu(In,Ga)Se2).
It is possible to connect PV cells in series to generate a sufficiently high voltage to
electrolyse water. There is an attraction, however, in developing a system that harvests
solar energy in a single cell to split the water molecules directly. The process is called
‘electrochemical photolysis’, or simply ‘photolysis’, and is effectively carried out in the
leaves of every living plant — the first step in the process known as ‘photosynthesis’. The
voltage required for water electrolysis of 1.293 V under standard conditions rules out
most metallic oxides and sulfides. A search has been ongoing for many years to identify
semiconductor materials that are able to exhibit at least the 1.6–1.7 V that is required
for water splitting under normal operating conditions.
In 1972, Fujishima and Honda were the first to demonstrate that hydrogen and oxygen
could be produced directly in a photoelectrochemical cell (PEC).28 Their cell used a
single‐crystal titanium dioxide (TiO2) electrode that was connected via a wire to a counter
electrode of platinum where hydrogen was released.
Unfortunately, titanium dioxide only absorbs light in the UV region, i.e., below a
wavelength of about 385 nm. To encourage absorption in the visible region, the dye‐
sensitized solar cell (DSSC) was developed. The fundamental operation involves the
absorption of a dye, usually ruthenium based, on the surface of the porous titania
electrode. When light strikes the dye, electrons are released into the conduction band
of the titania at the same time as the dye is oxidized. The electrons can then jump into
the valence band of the titania, which is intimately mixed with a current-collector that
is usually tin oxide coated on glass. The electrons flow around an external circuit, and
the dye is reduced via a ‘redox mediator’, which is typically the iodide–tri‐iodide couple
(I––I3–) dissolved in acetonitrile or some other organic solvent. The mediator diffuses
to the negatively charged electrode where it is reduced by electrons that have travelled
round the external circuit.
The DSSC can be used together with a conventional PV cell to produce a ‘tandem’ cell
in which a standard PEC is placed in front of a DSSC (Figure 10.17). The photoelectrode
in the PEC absorbs the high‐energy UV and blue light in sunlight to liberate oxygen,
while radiation of longer wavelengths in the green‐to‐red region of the spectrum passes
through and is absorbed by the DSSC. This boosts the flow of electrons that are fed back
to a counter electrode in the PEC to produce hydrogen. With such an arrangement,
photon efficiencies as high as 12% have been reported.
Since the discovery of the photochemical activity of titania, several other materials
have been found to exhibit the required activity, although much work remains to
produce candidates that have the required resistance to long‐term degradation. A
gadolinium semiconductor (Ga0.82Zn0.18)(N0.82O0.18) loaded with rhodium and chromium
photocatalyst was reported in 2008 to achieve a quantum yield in visible light of 5.9%.
28 Fujishima, AK and Honda, K, 1972, Electrochemical photolysis of water at a semiconductor electrode,
Nature (London) vol. 238, pp. 37–38.
313
314
Fuel Cell Systems Explained
e–
Dye on semiconductor
Platinum
electrode
Conducting
glass
Gas outlet
H2 O2
H2O
Violet
Conducting
glass
Green
Red
Aqueous
electrolyte
Optical
window
Semiconductor
Conducting
glass
Electrolyte
(mediator/organic solvent)
e–
Photoelectrochemical cell
Dye-sensitized solar cell
Figure 10.17 Operating principles of a tandem cell for enhanced hydrogen production.
Nevertheless, such materials display little absorption activity beyond about 440 nm and
thereby have an overall solar‐to‐hydrogen efficiency of only about 0.1%. In 2015, a new
low‐cost photoelectrochemical catalyst based on cobalt oxide was shown to have a
superior efficiency of around 5% and therefore may pave the way for other affordable
semiconductor nanomaterials.29
10.9 Thermochemical Hydrogen Production
and Chemical Looping
10.9.1 Thermochemical Cycles
Because of the stability of the water molecule and consequently the very high
temperatures that are required to split it thermally, attempts have been made to
accomplish the process at a more moderate upper temperature (i.e., <1000°C) by
means of an indirect route. The general idea is to decompose water by reacting it with
one or more chemicals that are regenerated via a series of cyclic thermochemical
29 Liao, L, Zhang, Q, Su, Z, Zhao, Z, Wang, Y, Li, T, Lu, X, Wei, D, Feng, G, Yu, Q, Cai, X, Zhao, J, Ren, Z,
Fang, H, Robles‐Hermandex, F, Baldelli, S and Bao, J, 2014, Efficient solar water‐splitting using a
nanocrystalline CoO photocatalyst, Nature Nanotechnology, vol. 9, pp. 60–73.
Fuels for Fuel Cells
reactions. In this way, the hydrogen and oxygen evolution reactions are separated.
Clearly, on practical and efficiency grounds, the fewer reactions involved the better.
Although the idea is sound from a thermodynamic point of view, there are both
engineering and materials issues.
The sulfur–iodine cycle has been studied extensively by the nuclear industry since the
mid‐1970s. In this cycle, iodine is employed to promote the oxidation of S(IV) to S(VI)
as follows:
I2 SO2 2H2O
2HI + H2SO 4
(10.21)
H2SO 4 → SO2 + H2O + ½O2 Endothermic : 850 − 900°C
(10.22)
2HI
(10.23)
I2 H2 Endothermic : 300 450 C
Two immiscible phases are first formed. The upper phase contains almost all of the
sulfuric acid, and the lower and dense phase holds most of the hydrogen iodide and
iodine. These are separated and the upper phase is then decomposed via reaction
(10.22), while the hydrogen iodide in the lower phase is converted into hydrogen and
iodine according to reaction (10.23). The products of the two reactions (SO2 and I2,
respectively) are then recycled to reaction (10.21).
In an attempt to reduce the number of steps in the thermochemical cycle from three
to two and at the same time to avoid the use of highly corrosive reagents such as sulfuric
acid, attention is on simpler cycles in which water is reduced to hydrogen by metals (M)
or metal oxides (MOred) in a lower oxidation state, i.e.,
M/MO red
H2O
MO ox +H2
(10.24)
In the second step of the cycle, the product oxide in its higher oxidation state is thermally
decomposed to liberate oxygen and revert to its original form, namely:
MO ox → M/MO red + ½O2
(10.25)
The classic example is that based on iron oxides as follows:
3FeO H2O
Fe3O 4
H2 Exothermic
Fe3O 4 (l ) → 3FeO(l ) + ½O2 Endothermic : > 1600°C
(10.26)
(10.27)
Note that the oxides are liquid at 1600°C. To avoid the high temperatures required by
reaction (10.27), the manganite (Fe3O4) can be replaced by the mixed oxide (Ni0.5Mn0.5)
Fe2O4, which is partially reduced to an oxygen‐deficient state. The first step in the
cycle can be carried out at about 800°C to regenerate the oxide and liberate hydrogen.
Even though the temperatures can be lowered, there are challenging issues in
engineering a reactor that can be cycled from oxidizing to reducing environments, as
well as in identifying material that does not degrade significantly over time with
repeated oxidation and reduction.
The many thermochemical cycles that have been investigated generally fall into four
groups, as shown in Table 10.7. Recent work is focused on low‐temperature reactions,
such as the copper chloride system.
315
316
Fuel Cell Systems Explained
Table 10.7 Thermochemical cycles currently under consideration.
Number of
reaction steps
Maximum
temperature (°C)
LHV
efficiency (%)
2
900
43
Sulfur cycles
Hybrid sulfur
(Westinghouse ISPRA, Mark 11)
(1150 without
catalyst)
Sulfur–iodine
3
900
(General Atomics, ISPRA Mark 16)
38
(1150 without
catalyst)
Volatile metal/oxide cycles
Zinc/zinc oxide
2
1800
45
Hybrid cadmium
2
1600
42
Non‐volatile metal oxide cycles
Iron oxide
2
2200
42
Cerium oxide
2
2000
68
Ferrites
2
1100–1800
43
4
530
39
Low‐temperature cycles
Hybrid copper chloride
A variation of the thermochemical cycle is to use a hydrocarbon gas as the means
of reducing the oxide back to the metal. An example would be replacing reaction
(10.24) with
CH 4 MO x
M/MO x CO2
2H2O
(10.28)
and specifically in the case of iron/iron oxide, this becomes
CH 4 Fe3O 4
3Fe CO2
2H2O
(10.29)
If the reduced iron is heated, it can be used to reduce steam to pure hydrogen, i.e.,
3Fe 4H2O
4H2 Fe3O 4
(10.30)
Equations (10.29) and (10.30) therefore provide the basis for an excellent means of
generating pure hydrogen from a hydrocarbon such as methane or natural gas.
Compared with the processes examined in this chapter for dealing with hydrocarbon
fuels, the use of a hydrocarbon gas offers the following three advantages:
●
●
●
Saving of investment costs associated with CO removal and other purification units
by cyclic operation of a single reactor.
Saving of investment and operational costs by using iron oxide as a cheap material.
High quality of the produced hydrogen.
One of the problems with the simple iron/iron‐oxide process is that the reduction
reaction (10.29) is very slow. Recent research has focused on improving the rate of this
Fuels for Fuel Cells
reaction, while maintaining moderate temperatures, by including other materials or
promoters either in combination with the iron or as a separate layer of catalyst in the
reactor. A mixture of ceria and zirconia has been shown to promote the rapid POX of
methane to CO and H2, which will then reduce the iron material according to:
4CO Fe3O 4
4CO2
3Fe
4H2 Fe3O 4
4H2O 3Fe
(10.31)
(10.32)
Even with such enhancements, there remains concern over the long‐term performance
and mechanical integrity of the iron/iron oxide.
10.9.2
Chemical Looping
Chemical looping combustion (CLC) is a process that is analogous to thermochemical
hydrogen production. Instead of performing hydrocarbon combustion in a single reaction
stage, two (or conceivably more) reactions are used. An additional species is required
and circulates between the two reactions. This additional species is typically a metal
and serves to carry oxygen between the reactions, that is, essentially the same function
as the iron/iron oxide described earlier. As an example, consider the following two
reactions, representing the CLC of methane, using a nickel-based reaction scheme:
4 Ni 2O2
CH 4
4 NiO
4 NiO
(10.33)
CO2 2H2O 4 Ni
(10.34)
If reactions (10.32) and (10.33) are added together, the nickel simply circulates between
the two reactions; hence, from the perspective of an overall mass and energy balance,
the two reactions simplify to the basic methane oxidation reaction, i.e.,
CH 4
2O2
CO2
2H 2 O
(10.35)
If both reactions (10.33) and (10.34) are arranged to take place in separate vessels,
the oxygen does not come into contact with the fuel. This gives an immediate advantage
when the source of oxygen is air, as will usually be the case. The result is that the
product gas is not diluted with nitrogen, i.e., the process incorporates a method of
separating CO2 from the combustion products (water is relatively easy to remove from
the products of reaction (10.34)). For this reason CLC has been investigated in detail
for use in industrial combustion processes as a means of separating CO2 prior to
sequestration.
Chemical looping combustion has also been proposed in conjunction with gasification
of hydrocarbons, including coal, for the production of hydrogen. Such a process may
include the following steps:
1) Generation of hydrogen from steam using suitable oxygen carriers.
2) Fuel gasification in the presence of an H2–steam mixture.
3) Combustion of the fuel off‐gas from the gasification process in the presence of
oxygen carriers.
4) Regeneration of oxygen carriers.
317
318
Fuel Cell Systems Explained
For the proposed system, the oxygen carrier is iron-based (Fe2O3/Fe3O4), and tests have
shown that the process offers advantages over conventional coal gasification with
CLC in terms of lower gasifier temperature (1068°C vs. 1700°C), increased hydrogen
production and improved cold gas efficiency.30 The process has the further dual
advantages of providing a ready means of CO2 separation prior to sequestration and the
elimination of an air separation unit for the gasifier. Again, the use of oxygen carrier
ensures that the hydrogen is produced with high purity and thereby is suitable for FCVs
and other applications of PEMFCs.
10.10
Biological Production of Hydrogen
10.10.1
Introduction
Reference has already been made in this chapter to the generation of hydrogen from
biofuels (see Section 10.3). Consideration was given to various biofuels and how they
could be used in fuel cells, mainly by converting to hydrogen‐rich gas via steam reforming
and other processes reviewed in Sections 10.4, 10.5 and 10.6. By contrast, this section
examines how biological methods might be employed to extract the hydrogen from any
fuel — natural gas as well as biofuels. These biological systems proceed through one of
three types of metabolic process as follows:
1) Photosynthesis by unicellular microorganisms that utilize either hydrogenase or
nitrogenase reactions.
2) Digestion through the action of bacteria to produce hydrogen anaerobically.
3) Various stepwise processes that use a combination of bacteria to predigest complex
organic molecules to make less complex organic material that can then be transformed with hydrogen‐producing organisms.
Progress in developing these processes has been slow due to the following issues:
●
●
●
●
The growth of organisms is inhibited by catabolites formed in microbiological
cultures.
Hydrogen production is limited because growth of organisms often slows down as
hydrogen concentrations build up.
There is only a narrow range of feedstocks on which the organisms can function.
The rate of production of other gases is high, or the rate of production of hydrogen is low.
Biological hydrogen generation is an active field of research, and, consequently, biologically
active enzymes are now able to be isolated and modified. Results are starting to emerge that
could have a profound influence on the perceptive of hydrogen as a future fuel.
10.10.2
Photosynthesis and Water Splitting
Photosynthesis consists of two processes: (i) the conversion of light energy to biochemical
energy by a photochemical reaction and (ii) the reduction of atmospheric carbon dioxide
30 Zhang, Y, Doroodchi, E and Moghtaderi, B, 2012, Thermodynamic assessment of a novel concept for
integrated gasification chemical looping combustion of solid fuels, Energy & Fuels, vol. 26, pp. 287–295.
Fuels for Fuel Cells
to organic compounds such as sugars. In the first process, light is absorbed by chlorophyll,
which acts as a mediator in the oxidation of water:
(10.36)
2H2O + 2h
4H + 4e + O2
In the second process, the organic compound nicotinamide adenine dinucleotide
phosphate (NADP) is reduced by electrons to a state generally designated as NADPH.
Along with adenosine 5′‐triphosphate (ATP), NADPH is an important intermediary
in the photosynthetic fixation of carbon dioxide. The protons and electrons react with
carbon dioxide via these two mediators to produce sugars. The overall photo biochemical
process taking place in green plants is represented by:
nCO2 + 2nH2O ATP NADPH
n(CH2O) nH2 O nO2
(10.37)
where n is defined according to the structure of the resulting carbohydrate.
Like the chlorophyll in plants, the pigments in some types of algae can absorb solar
energy under certain conditions. A few groups of algae and cyanobacteria (formerly
known as ‘blue‐green algae’) produce gaseous hydrogen rather than sugars via a
photosynthesis route. Cyanobacteria contain hydrogenase or nitrogenase enzymes, and
it is these that have the ability to catalyse hydrogen formation.
In 1942 it was observed that the green alga Scenedesmus produces hydrogen when
exposed to light after being kept in the dark and under anaerobic conditions.31 Further
work directed towards elucidating the mechanism of this process identified hydrogenase
as the key enzyme, which reduces water to hydrogen with concomitant oxidation of an
electron carrier, ferredoxin. The green algae therefore became known as ‘water‐splitting’
organisms, and the process conducted in the photo biochemical cells was referred to as
‘biophotolysis’. Unfortunately, the hydrogenase in green algae is very sensitive to oxygen,
which can rapidly deactivate the enzyme’s activity. In 2007, researchers found that production of oxygen by the hydrogenase could be blocked by adding copper to the algae.
Sulfur had a similar effect. At the same time the Solar Biofuels Consortium — a collaboration between scientists at the University of Bielefeld (Germany) and Queensland
(Australia) — managed to genetically modify the single‐cell green alga Chlamydomonas
reinhardtii in such a way that it produces an especially large amount of hydrogen. The
work demonstrated production rates up to five times the volume made by the wild form
of alga and up to 1.6–2.0% energy efficiency.32
Although other species have exhibited good hydrogen production activity, e.g.,
Chlamydomonas moewusii and Anabaena cylindrica, C. reinhardtii has received the
most attention from researchers, and scale‐up of algal bioreactors is now in progress.
The unicellular aerobic nitrogen fixer Synechococcus sp. Miami BG043511 is
another biological system that exhibits photosynthesis activity. It has provided a
conversion efficiency estimated at around 3.5% based on photosynthetically active
radiation (PAR) (i.e., light of energy 400–700 nm in wavelength) and using an artificial
31 Gaffron H. and Rubin J, 1942, Fermentative and photochemical production of hydrogen in algae, Journal
of General Physiology, vol. 26, pp. 219–240.
32 Hankamer, B, Lehr, F, Rupprecht, J, Mussgnug, JH, Posten, C and Kruse, O, 2007, Photosynthetic biomass
and H2 production by green algae: from bioengineering to bioreactor scale up. Physiologia Plantarum,
vol. 131, pp. 10–21.
319
320
Fuel Cell Systems Explained
light source. Certain other bacteria are also able to photosynthesize, but not via water
oxidation. These function with either organic compounds or reduced sulfur compounds
as the electron donors. The conversion efficiency of light energy to hydrogen in such
systems can be much higher than that achieved with cyanobacteria. For example,
efficiencies of 6–8% have been exhibited by Rhodobacter sp. in the laboratory. It is
likely that solar efficiencies of 10% will soon be achieved by such bacteria.
10.10.3
Biological Shift Reaction
In 1997, it was demonstrated that certain bacteria (Rhodospirillum rubrum) can utilize
carbon monoxide and water to produce carbon dioxide and hydrogen gas via the WGS
reaction. Subsequently, the US DOE has funded the research and development of this
type of biological process. The implications of carrying out the WGS reaction in a bioreactor are important since normally catalysts operating at high temperatures are
required to ensure reasonable rates. If the reaction could be carried out at near-ambient
temperature, the product gas would contain little CO and therefore further expensive
gas clean‐up for PEMFCs could be eliminated or at least greatly reduced.
In a recent study, the WGS reaction was separated into two half‐cell electrochemical
reactions, namely, H+ reduction and CO oxidation. The former reaction was catalysed
by a hydrogenase, Hyd‐2, from Escherichia coli, and the latter reaction by a carbon
monoxide dehydrogenase (CODH I) from Carboxydothermus hydrogenoformans. Both
enzymes were attached to conducting carbon particles. The resulting electrocatalyst
proved to be highly active for WGS when compared with more conventional, high‐
temperature, supported‐metal WGS catalysts.33
10.10.4
Digestion Processes
Hydrogen can be produced by microbial digestion of organic matter in the absence
of light energy. Under relatively mild conditions of temperature and pressure, many
bacteria will readily produce hydrogen together with acetic acid and other low molecular
weight organic acids. Nevertheless, the rates of reaction are usually low, and hydrogen
is not produced in significant amounts because of two mitigating factors. First,
inhibition of the microbial hydrogenase may occur as hydrogen builds up. Second,
hydrogen may react with other organic species that may also be present, or with carbon
dioxide in the system, to cause the generation of methane. As the partial pressure of
hydrogen increases, the forward reaction of organic matter to hydrogen becomes
thermodynamically unfavourable. The challenge in using digestion processes is
therefore to increase the rate of hydrogen production while preventing methane
formation. Hydrogen production generally occurs by fermentation of carbohydrate‐
rich material in the organic waste and is carried out by anaerobic bacteria belonging
to species such as Enterobacter, Bacillus and Clostridium. More recently, the
thermophilic Thermotoga neapolitana has also demonstrated considerable promise.
33 Lazarus, O, Woolerton, TW, Parkin, A, Lukey, MJ, Reisner, E, Seravalli, J, Pierce, E, Ragsdale, SW, Sargent,
F and Armstrong, FA, 2009, Water-gas shift reaction catalyzed by redox enzymes on conducting graphite
platelets, Journal of the American Chemical Society, vol. 131(40), pp. 14154–14155.
Fuels for Fuel Cells
This bacterium has the potential to utilize a variety of organic wastes and to offer a
cost‐effective method of generating significant quantities of hydrogen.
Microbial digestion in the absence of light (‘dark fermentation’) produces a mixed
biogas containing primarily hydrogen and carbon dioxide. To maximize hydrogen
production, it is necessary to optimize the activity of the enzyme hydrogenase. In this
respect, recent studies have shown that the pH should be maintained in the range of
5–6.5, with an optimum value of 5.5.34 Digestion is attractive for treating sewage sludge
as well as agricultural waste that would otherwise have low value, such as cheese whey
and dairy manure. Unfortunately, the yield of hydrogen (e.g., as compared with methane)
from anaerobic digestion is low.
Although biological processes for hydrogen production are actively being developed,
many questions remain concerning the fundamental biochemical processes that are
occurring. Photosynthesis‐algae or photosynthesis‐bacteria systems seem to be the best
candidates for the first technical applications. Present indications are that hydrogen
production costs of 12 cents per kWh H2 or less are achievable.
Further Reading
Brown, RC and Stevens, C, 2011, Thermochemical Processing of Biomass: Conversion into
Fuels, Chemicals and Power, John Wiley & Sons, Inc., Hoboken, NJ. ISBN:978‐0‐470‐
72111‐7.
Carmo, M, Fritz, DL, Mergel, J and Stolten, D, 2013, A comprehensive review on PEM
water electrolysis, International Journal of Hydrogen Energy, vol. 38(12), pp. 4901–4934.
Dincer, I and Joshi, AS, 2013, Solar Based Hydrogen Production Systems, Springer,
New York. DOI 10.1007/978‐1‐4614‐7431‐9_2. ISBN 978‐1‐4614‐7430‐2.
Hallenbeck, PC (ed.), 2012, Microbial Technologies in Advanced Biofuels Production, 15,
Springer US, Boston, MA. DOI 10.1007/978‐1‐4614‐1208‐3_2.
Hoogers, G, 2003, Fuel Cell Technology Handbook, CRC Press, Boca Raton, FL. ISBN
0‐8493‐0877‐1.
HTGR‐integrated hydrogen production via steam methane reforming (SMR) process
analysis, 2010, Technical Evaluation Study Project No. 23843, Idaho National
Laboratory.
Kahn, MR, 2011, Advances in Clean Hydrocarbon Fuel Processing: Science and Technology,
Series in Energy, Woodhead Publishing, Philadelphia, PA. ISBN‐10: 1845697278.
Kidnay, AJ, Parrish, WR and McCartney, DG, 2010, Fundamentals of Natural Gas
Processing, 2nd edition, CRC Press, Boca Raton, FL. ISBN‐13:978‐1420085198.
Kolb, G, 2010, Fuel Processing: For Fuel Cells, Wiley‐VCH, Weinheim. ISBN:
978‐3‐527‐31581‐9.
Rand, DAJ and Dell, RM, 2008, Hydrogen Energy: Challenges and Prospects, The Royal
Society of Chemistry, Cambridge. ISBN: 978‐0‐85404‐597‐6.
34 Valdez–Vazquez, I and Poggi–Varaldo, HM, 2009, Hydrogen production by fermentative consortia,
Renewable and Sustainable Energy Reviews, vol. 13, pp. 1000–1113.
321
323
11
Hydrogen Storage
11.1
Strategic Considerations
Society has readily adapted to many different fuels such as petroleum, diesel oil,
naphtha, coal and biofuels (e.g., synthetic diesel, ethanol–gasoline blends) — but
hydrogen is different. As a gas under normal temperature and pressure, hydrogen
presents its own challenges as a fuel. It has to be generated either by extraction from
another fuel or by the decomposition of water by electrolysis or photo‐electrolysis. Unlike
electricity, however, that has become ubiquitous and the backbone of conventional
energy‐supply networks, hydrogen can be stored relatively easily in bulk. There should be
no surprise therefore that hydrogen has been promoted in recent years as a prospective
means of storing renewable energy through offering long‐term advantages over battery
systems. Over the past decade, the issue of climate change has made the case for zero‐
emission vehicles so strong that hydrogen has almost universally been adopted as the
fuel of choice for fuel‐cell vehicles (FCVs). This is remarkable considering the substantial
technical, environmental and fiscal challenges associated with the development of
hydrogen infrastructure. A summary of some applications for which hydrogen is already
finding use, together with current example systems, is presented in Table 11.1.
Small fuel‐cell systems of a few kW or below are generally designed to run on stored
hydrogen because it is difficult to scale down conventional methods of hydrogen
generation. Whereas for systems of about 50 kW and greater it is cost‐effective to couple
the fuel processor directly with the fuel‐cell stack, this is not the case for small proton‐
exchange membrane fuel cell (PEMFC) systems. A small local store of hydrogen is
therefore an essential component of fuel‐cell systems for portable applications, unless
the direct methanol fuel cell (DMFC) is being employed.
Hydrogen can also be a reasonable way of storing electrical energy from sources such
as wind‐driven generators and hydroelectric power, where generation might well be out
of line with consumption. Electrolysers convert the electrical energy to hydrogen during
times of high supply and low demand. In general, the storage of hydrogen for stationary
applications is less demanding than for transportation systems in which there are
more severe constraints in terms of acceptable mass and volume, speed of charge
and discharge and, for some storage systems, heat management. Finding a satisfactory
solution for the on‐board containment of hydrogen has proven to be a major challenge
in the development of FCVs.
Fuel Cell Systems Explained, Third Edition. Andrew L. Dicks and David A. J. Rand.
© 2018 John Wiley & Sons Ltd. Published 2018 by John Wiley & Sons Ltd.
324
Fuel Cell Systems Explained
Table 11.1 Applications that employ hydrogen storage.
Application
Method of hydrogen storage
Example system
Fuel‐cell car
Compressed gas, 70 MPa
Toyota Mirai, 2015
Hydrogen internal
combustion‐engined car
Liquid hydrogen, −252°C
BMW series 7 sedan
Fuel‐cell bus
Compressed gas, 70 MPa
Mercedes‐Benz Citaro
Fuel‐cell bicycle/
motorcycle
Compressed gas
Palcan, Intelligent Energy
Locomotive
Low‐temperature hydride
Anglo‐American Platinum, mining
locomotive, Dishaba Mine, South
Africa
Airplane
Liquid hydrogen
Boeing Phantom‐Eye
Fuel‐cell battery charger
Low‐temperature hydride
Intelligent Energy, Upp, and
Horizon MiniPak portable
chargers
Renewable energy
storage
Low‐temperature hydride
Sir Samuel Griffith Centre
(30 MWh), Australia
Compressed gas, 70 MPa
INGRID project (39 MWh), Italy
Natural gas + hydrogen
blend
Compressed gas
ENEA, Regione Emilia Romagna,
Italy, Hydrogen‐Compressed
Natural Gas (HCNG) bus trials
Industrial gas
distribution
Medium‐temperature hydride
Hydrexia, Ni–Mg hydride
Spacecraft
Liquid hydrogen
Government space agencies of
China, Europe, India, Russia and
the United States
McPhy, Mg hydrogen
Although hydrogen has very high specific energy (Wh kg−1), which makes it the fuel of
choice for space missions, its volumetric energy density (Wh m−3) is very low. The latter
feature is a great disadvantage compared with most other fuels. The energy density can
be improved by compressing the gas, and pressures up to 70 MPa are routinely used for
storage on FCVs. Unlike liquefied petroleum gas (LPG) or butane, which can be liquefied
at ambient temperature by raising the pressure, hydrogen can only be liquefied by cooling
the gas down to about 22 K. As a liquid, the energy density is quite low, 71 kg m−3. It has
to be stored in a thermally insulating ‘Dewar’ vessel, but even with the best designs
some loss (‘boil‐off ’) of liquid hydrogen by evaporation is inevitable.
Hydrogen can also be stored in various chemical compounds that, on a gravimetric
basis, can hold quite large quantities of hydrogen. To be useful, a given compound must
pass the following three tests:
1) The compound must readily give up the hydrogen — otherwise there is no advantage
over using a reformed fuel in one of the ways described previously in Chapter 10.
2) The manufacturing process must be simple and absorb little energy — in other words
the energy and financial costs of hydrogen becoming incorporated in the compound
must be low.
3) The compound must be safe to handle.
Hydrogen Storage
Table 11.2 Possible materials for hydrogen storage.
Formula
Percent
hydrogen
Liquid hydrogen
H2
100
0.07
14.0
Cold, −252°C
Liquid methane
CH4
25.13
0.422
9.6
Cold, −175°C
Ammonia
NH4
17.76
0.682
8.5
Toxic,
>100 ppm
Name
Density
(kg L−1)
Volume (L) to
store 1 kg H2a
Notes
Liquids/gases
Water
H2O
11.11
1.00
8.9
Hydrazine
N2H4
12.58
1.011
7.8
Methanol
CH3OH
12.50
0.79
10.0
Ethanol
C2H5OH
13.00
0.790
9.7
6.30
1.060
15.0
30 wt.% sodium
NaHBH3 + H2O
borohydride solution
Toxic, >10 ppm
Expensive, but
works well
Simple hydrides
Lithium hydride
LiH
12.68
0.82
6.50
Sodium hydride
NaH
4.30
0.92
25.9
Diborane
B2H6
21.86
0.417
11.0
Beryllium hydride
BeH2
18.28
0.67
8.2
Silane
SiH4
12.55
0.68
12.0
Calcium hydride
CaH2
5.00
1.90
11.0
Aluminium hydride
AlH3
1.30
7.1
Potassium hydride
KH
2.51
1.47
27.1
Titanium hydride
TiH2
4.40
3.90
5.8
10.8
Caustic
Caustic but
cheap
Toxic
Highly toxic
Toxic, >5 ppm
Caustic
Complex hydrides
Lithium borohydride
LiBH4
18.51
0.666
8.1
Mildly toxic
Aluminium
borohydride
Al(BH4)3
16.91
0.545
11.0
Mildly toxic
Lithium aluminium
hydride
LiAlH4
10.62
0.917
10.0
Palladium hydride
Pd2H
Titanium–iron hydride TiFeH2
a)
0.47
10.78
20.0
1.87
5.47
9.8
All the extra equipment required to hold or process the compound is excluded. Consequently, each
entry is not a total number and should only serve as a guide. For example, all the alkali metal hydrides
require large quantities of water, from which some of the hydrogen is also released.
A large number of promising chemicals have been suggested or tested; some examples,
together with their key properties, are listed in Table 11.2. Unfortunately, many of the
candidates do not warrant a great deal of consideration, as they fail one or more of the
three tests. Hydrazine, for instance, passes the first test very well (it has been successfully
demonstrated in fuel‐cell systems — see Section 5.2.3, Chapter 5) but is highly toxic and
325
326
Fuel Cell Systems Explained
very energy intensive to manufacture, and so it fails the second and third tests.
Nevertheless, several other hydrogen compounds have found practical application and
are described in detail later on. The most important of these involves the use of metal
hydrides. There are two principal forms: a ‘rare earth’ metal hydride compound that can
reversibly store and deliver hydrogen and alkali metal hydrides that react with water to
yield hydrogen gas.
The principal methods of storing hydrogen that will be described in this chapter are
therefore:
Compression in gas cylinders.
As a cryogenic liquid.
As a reversible metal hydride.
As a metal hydride that reacts with water.
●
●
●
●
Before discussing these methods, which each have advantages and disadvantages,
consideration must be given to issues of safety.
11.2
Safety
Compared with all gases, hydrogen has (i) the lowest molecular weight, (ii) the highest
thermal conductivity, velocity of sound and mean molecular velocity and (iii) the lowest
viscosity and density. Accordingly, hydrogen leaks through small orifices faster than all
other gases, e.g., 2.8 and 3.3 times faster than methane and air, respectively. In addition,
hydrogen is a highly volatile and flammable gas, and in certain circumstances mixtures
of hydrogen and air can detonate. Safety considerations must therefore feature strongly
in the design of any fuel‐cell system. All measures should be taken to avoid the risk of
hydrogen escape, and systems should be equipped with sensors to shut off gas supplies
and alert personnel should leaks occur.
Although safety should be a priority, it is important to stress that hydrogen is no
more dangerous, and in some respects it is rather less dangerous, than various other
conventional fuels. The key properties relevant to the safety of hydrogen and two
other gaseous fuels that are widely available — methane and propane — are listed in
Table 11.3. The lower concentration limit for ignition of hydrogen is much the same
Table 11.3 Properties relevant to safety for hydrogen and two other commonly used gaseous fuels.
Density (kg m−3 at NTPa)
a
Hydrogen
Methane
Propane
0.084
0.65
2.01
Ignition limits in air (vol.% at NTP )
4.0–77
4.4–16.5
1.7–10.9
Ignition temperature (°C)
560
540
487
Minimum ignition energy in air (MJ)
0.02
0.3
0.26
Maximum combustion rate in air (m s−1)
3.46
0.43
0.47
Detonation limits in air (vol.%)
18–59
6.3–14
1.1–1.3
Stoichiometric ratio in air
29.5
9.5
4.0
a)
NTP, normal temperature and pressure is 20°C and 101.325 kPa.
Hydrogen Storage
as for methane. For propane a lower concentration is necessary for ignition. The
ignition temperatures for hydrogen and methane are similar, but both are higher
than that for propane. The minimum ignition energy of hydrogen is, however, very
low, and thereby it suggests that a fire could be started very easily. In fact, the ignition
energies of all the gases are actually lower than would be encountered in most
practical cases. A spark could ignite any of these three fuels. Furthermore, against
this must be set the much higher minimum concentration needed for detonation of
hydrogen in air.
Another potential hazard arises from the rather greater range of concentrations
required to cause detonation of hydrogen. Care must therefore be taken to prevent the
accumulation of hydrogen in confined spaces. Fortunately, the high buoyancy and high
average molecular velocity mean that of all common gases, hydrogen disperses most
rapidly and is therefore less likely than other gases to build up to a level required for
detonation.
Comparing the data given in Table 11.3, hydrogen appears much the same as the
other fuels from the point of view of potential danger. It is the much lower density that
gives hydrogen an advantage from a safety point of view. The density of methane is
similar to air, which means it does not disperse quickly but tends to mix with air.
Propane has a lower density than air, which tends to make it sink and collect at low
points, such as in basements, drains and the hulls of boats, where it can explode or be
set alight with devastating effects. Hydrogen, on the other hand, is so light that it rapidly
disperses upwards. The concentration levels necessary for ignition or detonation are
therefore much less likely to be achieved for hydrogen in comparison with the other fuels.
Hydrogen, like all fuels, must be handled carefully. All things considered, however,
it does not present any greater hazard than any other flammable liquids or gases
encountered today. The one unique characteristic that should always be borne in
mind by developers or users of fuel cells is that, once ignited, hydrogen burns with an
invisible flame.
11.3
Compressed Hydrogen
11.3.1
Storage Cylinders
When hydrogen is produced at a central facility, it can either be stored in bulk before
dispatch to customers or first distributed and then stored locally on‐site until needed.
Whatever form delivery is taken, e.g., as gas in cylinders or in pipelines or as liquid
hydrogen, it is first necessary to compress the gas. To do this, work has to be done,
and energy expended. Broadly, the energy required to compress hydrogen lies
between 5 and 15% of the higher heating value (HHV), as determined by the final
pressure, and whether the process is carried out adiabatically or isothermally. In
practice, a multistage compression process will probably be adopted, and if mechanical
and electrical losses are included, the total energy wasted in compression can amount
to 20%.
The main advantages of storing hydrogen as a compressed gas are:
●
●
●
Simplicity.
Indefinite storage time.
No purity limits on the hydrogen.
327
328
Fuel Cell Systems Explained
Storing hydrogen gas at pressure in steel cylinders is the most technically straightforward method of holding hydrogen and the most widely adopted for small amounts
of gas. Tube trailers, in which the compressed gas is contained in horizontal steel
cylinders, are the normal delivery method for merchant hydrogen, i.e., where the
quantities involved are considerable. The steel cylinders are permanently fixed to
the trailer and are discharged in situ, i.e., not offloaded; a trailer is shown in
Figure 11.1.
Small‐scale consumers, such as laboratories, employ cylinders with storage capacities
of just a few cubic metres (pressurized to 20 MPa) that can be manhandled with simple
trolleys. The specifications of one such cylinder are compared in Table 11.4 with those
of a larger cylinder used for storage on a bus or other road vehicle. The latter is
constructed with an aluminium inner liner of 6 mm thickness, around which is wrapped
a composite of aramid fibre and epoxy resin. This material has a high ductility and
thereby has good burst behaviour in that it rips apart rather than disintegrating into
many pieces. The burst pressure is 120 MPa.1
Figure 11.1 Tube trailer for delivery of compressed hydrogen gas — 30 tubes and 1225 m3 (105 kg)
capacity. (Source: Reproduced with permission of Coregas.)
1 It should be noted, however, that at present composite cylinders are about three times the cost of steel
cylinders of the same capacity.
Hydrogen Storage
Table 11.4 Comparative data for two cylinders used to store hydrogen at high pressure.
2‐L steel (20 MPa)
147‐L composite (30 MPa)
Mass of empty cylinder (kg)
3.0
100
Mass of hydrogen stored (kg)
0.036
3.1
Storage efficiency (wt.% H2)
1.2
3.1
Specific energy (kWh kg−1)
0.47
1.2
a
3
Volume of tank (approximately), L (m )
2.2 (0.0022)
Mass of H2 (kg L−1)
0.016
a)
220 (0.22)
0.014
The storage efficiency here is defined as the total mass of hydrogen stored divided by the mass of the
empty cylinder, expressed as wt.% H2.
11.3.2
Storage Efficiency
In considering the storage of hydrogen on an FCV, it would be expected that a larger
storage system would be a great deal more efficient than a smaller one in terms of
specific mass of hydrogen stored. Therefore, the larger vessel (147 L) in Table 11.4
would be expected to hold proportionately more hydrogen than the (2 L) vessel. It
should be remembered, however, that large tanks have to be secured in the vehicle,
and consequently the weight of the supporting structure should be taken into
account in assessing efficiency, which is defined as the mass of hydrogen stored
divided by the mass of the storage medium (in this case the empty vessel). In one of
the early European buses, in which hydrogen was used to fuel an internal combustion
engine (ICE), 13 composite tanks were mounted in the roof space. The total mass of
the tanks and the bus structure reinforcements was 2550 kg, or 196 kg per tank. This
should be compared with the mass of an empty composite tank given in Table 11.4
of 100 kg. The increase in mass from 100 to 196 kg has the effect of reducing the
‘storage efficiency’ of this particular system to 1.6 wt.%, i.e., not so very different
from that of the (2 L) steel cylinder listed in Table 11.4. Another point is that the
weight of connecting valves, pressure‐reducing regulators and other essential hardware
has also to be taken into account whether the system is built around steel or
composite cylinders. When employing a 2.2 L steel cylinder, these components
would typically add about 2.15 kg to the mass of the system and thus reduce the
storage efficiency from 1.2 to 0.7 wt.%.2 The low density of hydrogen is responsible
for the poor mass storage efficiency, even at such high pressures. The density of
hydrogen gas at ambient temperature and pressure is 0.084 kg m−3, whereas air is
about 1.2 kg m−3. In practice, compressed hydrogen often accounts for less than 2%
of the total mass of the storage system.
2 Kahrom, H, 1999, Clean hydrogen for portable fuel cells, Proceedings of the European Fuel Cell Forum
Portable Fuel Cells Conference, 21–24 June 1999, Lucerne, pp. 159–170.
329
330
Fuel Cell Systems Explained
11.3.3
Costs of Stored Hydrogen
For small‐scale fuel cells, by taking into account all expenditure, e.g., depreciation of
cylinders, administration and purchase of pressure‐reducing valves, it has been
estimated that the cost of hydrogen fuel is about US$2.2 per g2. Using data given in
Section A2.4, Appendix 2, this expenditure corresponds to about US$56 per kWh, or
about US$125 per kWh for the electricity from a fuel cell of 45% efficiency. This is
absurdly expensive when compared with mains electricity but is noticeably cheaper
than current battery storage.3
11.3.4
Safety Aspects
The metal employed in the fabrication of the hydrogen storage vessel requires careful
selection. Hydrogen is a very small molecule and is capable of diffusing into materials
that are impermeable to other gases. This is compounded by the fact that a very small
fraction of the hydrogen gas molecules may dissociate on the surface of the material.
Diffusion of atomic hydrogen into the material may then occur and compromise its
mechanical integrity. Gaseous hydrogen can build up to form internal blisters in the
material that, in turn, can lead to crack promotion. With carbonaceous metals, such as
steel, the hydrogen can react with the carbon to produce bubbles of entrapped methane.
The gas pressure in the resulting internal voids can generate an internal stress that is
sufficient to cause fissures, cracks or blisters in the steel. The phenomenon is well
known and is termed ‘hydrogen embrittlement’. Fortunately certain chromium‐rich
steels and chromium–molybdenum alloys are resistant to hydrogen embrittlement, and
composite reinforced plastic materials can also be used for larger tanks, as mentioned
previously.
As well as the problem of the very high mass associated with the storage vessel, there
are significant hazards associated with storing hydrogen at high pressure. A leak from
such a cylinder would generate very large forces as the gas is released. It is possible
for such cylinders to become essentially jet‐propelled torpedoes and thereby inflict
considerable damage. Furthermore, vessel fracture would most likely be accompanied
by auto‐ignition of the released hydrogen in air and give rise to a fire that lasts until the
contents of the ruptured, or accidentally opened, vessel are consumed. Nevertheless,
compressed hydrogen is widely and safely used by correctly following established
procedures and guidelines. In vehicles, for example, pressure‐relief valves or rupture
discs (see following Section 11.4) are fitted and will safely vent gas in the event of a fire.
Similarly, pressure regulators attached to hydrogen cylinders are equipped with flame
traps to prevent ignition of the gas.
Compressed hydrogen storage is most widely employed in places where there is
moderate but variable demand. It is also used for road vehicles, both those with ICEs
and those with fuel‐cell systems. Figure 11.2 shows an example of composite compressed
hydrogen storage tanks on an FCV.
3 Lead–acid and lithium‐ion batteries cost around US$250 and US$300 per kWh, respectively, in 2015.
Nykvist, B and Nilsson, M, 2015, Rapidly falling costs of battery packs for electric vehicles, Nature Climate
Change, vol. 5, pp. 329–332.
Hydrogen Storage
Figure 11.2 Composite hydrogen storage tanks on the Honda FCX fuel‐cell car chassis. (Source:
https://zh.wikipedia.org/wiki/%E6%B0%AB%E6%B0%A3%E7%AE%B1. CC BY‐SA 3.0.)
11.4
Liquid Hydrogen
The storage of hydrogen as a liquid (denoted as LH2) at about 20 K is currently the only
method that is widely practised for storing substantial quantities of hydrogen. A gas
cooled to the liquid state in this way is known as a ‘cryogenic liquid’, and large amounts
are currently required for processes such as petroleum refining and ammonia production.
NASA is another notable customer and has huge 3200 m3 (850 000 US gallon) tanks to
ensure a continuous supply for space exploration (see Figure 11.3a).
The container or tank for storing cryogenic liquid hydrogen is a large, strongly
reinforced vacuum (Dewar) flask. Given that the container is never perfectly insulated,
liquid hydrogen will slowly evaporate and cause the pressure in the container to rise.
The maximum allowable pressure is normally below 300 kPa, though some larger
tanks may be designed for greater pressures. If the rate of evaporation exceeds the
demand, then the tank is occasionally vented to make sure the pressure does not rise
too high. A spring‐loaded valve will release and then close again when the pressure
falls. The small amounts of hydrogen involved are routinely released to the atmosphere,
though in very large systems they may be vented out through a flare stack and burnt.
A rupture disc is commonly fitted as a backup safety feature. This device consists of a
ring covered with a membrane of controlled thickness that will withstand a certain
safe pressure. When the pressure in the system exceeds the allowable pressure of the
rupture disc, the membrane bursts, and gas is released and safely vented away until the
disc is replaced. Usually, the fault cannot be rectified until all of the gas has been
vented from the storage system.
When an LH2 tank is being filled, and when fuel is being withdrawn, it is most
important that air is not allowed into the system; otherwise, an explosive mixture
331
332
Fuel Cell Systems Explained
(a)
(b)
(c)
(d)
Figure 11.3 Examples of cryogenic LH2 tanks: (a) NASA bulk storage for space missions. (b) installation
for AC Transit fuelling station at Emeryville, California. (c) Tank in the rear of BMW series 7 car.
(d) Proposed bulk ocean transport from Australia to Japan by Kawasaki Heavy Industries. (Source:
Reproduced with permission of Kawasaki Heavy Industries.)
could form. The tank therefore should be purged with nitrogen before filling.
Figure 11.3b shows an example of an LH2 tank installed as part of vehicle filling station
for AC Transit in California.
Considerable effort has gone into the design and development of LH2 tanks for
cars, though not specifically for FCVs. Several automobile companies have invested in
hydrogen‐fuelled ICEs, most of which have employed LH2. The tank in the BMW
hydrogen cars, for example (see Figure 11.3c), is cylindrical in shape and has the
conventional double‐wall vacuum flask type of construction. The walls are about 3 cm
thick and consist of 70 layers of aluminium foil interlaced with fibreglass matting. The
maximum operating pressure is 500 kPa. The tank stores 120 L of cryogenic hydrogen
that, as the density of LH2 is very low (~71 kg m−3), weighs only 8.5 kg. The key features
of the BMW cryogenic tank are given in Table 11.5.
One of the problems associated with cryogenic hydrogen is that the liquefaction
process is very energy intensive. Several stages are required. After an initial stage of
compression, the gas is cooled to about 78 K under the action of liquid nitrogen. The
hydrogen then is cooled further by expansion through a turbine. Finally, a magnetocaloric
Hydrogen Storage
Table 11.5 Details of a cryogenic hydrogen container
for BMW cars.
Mass of empty container (kg)
51.5
Mass of hydrogen stored (kg)
8.5
Storage efficiency (% mass H2)
14.2
Specific energy (kWh kg−1)
5.57
3
Volume of tank (approximately, m )
0.2
Mass of H2 (kg L−1)
0.0425
process is performed to convert ortho‐H2 to para‐H2.4 The total energy required to
liquefy the gas is about 40% of the specific heating value of the hydrogen.
In addition to the regular safety concerns with hydrogen, there are a number of specific
difficulties concerned with cryogenic storage. All pipes containing the fluid must be
insulated, as must any parts in good thermal contact with these pipes. This precaution is
necessary to minimize the chance of frostbite if human skin is brought into contact with
the system. Insulation is also necessary to prevent the surrounding air from condensing
on the pipes; otherwise an explosion could develop if liquid air drips onto nearby
combustibles. Asphalt, for example, can ignite in the presence of liquid air. Thus, concrete
paving is laid around static installations. In general, however, the hazards of hydrogen
are somewhat less with LH2 than with pressurized gas. For instance, if there is a failure of
the container, the fuel tends to remain in place and vent to the atmosphere more slowly.
Certainly, LH2 tanks have been approved for use in cars in Europe. Kawasaki has designed
an ocean‐going tanker to transport hydrogen from Australia to Japan (the concept is
shown in Figure 11.3d).
Filling stations for fuel‐cell vehicles are starting to become well established in numerous
locations — the United States, Japan and Europe. Most of these incorporate high‐pressure
storage and some also store liquid hydrogen. Figure 11.4 shows an example of a filling station
in the United Kingdom that delivers compressed hydrogen generated by electrolysis.
11.5
Reversible Metal Hydrides
Certain metals and alloys have the ability to absorb and release gaseous hydrogen,
reversibly, via the formation of hydrides, i.e.,
M
x
H2
2
MH x heat
(11.1)
The hydrogen molecule first dissociates into its two atoms, which are chemisorbed on
the surface of the metal/alloy, and then diffuses into the bulk lattice. The dissolved
atoms can take the form of a random solid solution or react to produce a hydride of
fixed stoichiometric composition. The quantity of hydrogen absorbed is expressed in
terms of the hydride composition, either on a molar basis (MHx) or on a wt.% basis.
4 Both atoms in the molecule of ortho‐hydrogen have nuclear spins in parallel. In para‐hydrogen, the spins
are antiparallel. The process of converting ortho‐ to para‐hydrogen is exothermic and if allowed to take
place naturally would cause boil‐off of the liquid.
333
334
Fuel Cell Systems Explained
Figure 11.4 Refilling with H2. (Source: http://theconversation.com/hydrogen‐car‐progress‐hasnt‐
stalled‐yet‐15821. CC BY‐ND 4.0.)
The rates of absorption and desorption can be controlled by adjusting the temperature
or pressure. (Note: Heat is liberated during formation of the hydride and therefore must
be added to effect its subsequent decomposition with discharge of the hydrogen.)
Hydrides can be tailored to operate over a wide range of temperatures and pressures. To
be most useful as a hydrogen store, the metal or alloy should react with the gas at, or
near, ambient temperature and at not too great a pressure. The enthalpy of absorption
should also not be too high. Otherwise, heat transfer becomes a problem, especially in
large hydride beds. Finally, the system should be capable of sustaining a practical number
of absorption–desorption cycles without deterioration.
The remarkable hydriding properties of LaNi5 were discovered in about 1969 at the
Philips laboratories in the Netherlands. It has served as a benchmark for many hydriding
alloys to be developed by other researchers. Metal alloys that have been designed and
developed in succeeding years generally fall into one of the following types: AB5, A2B7,
AB3, AB2 (Laves phase), AB and A2B. In these alloys, A represents a metal element with
a strong affinity for hydrogen (i.e., an ability to absorb hydrogen), and B is a metallic
element with weak hydrogen affinity but strong activity for catalysing the dissociation
of H2 molecules to H atoms. Of these classes of materials, the most studied are probably
the AB5 alloys in which A is calcium or a rare earth element and B is customarily a
transition metal (e.g., Fe, Co, Ni, V, Mn, Cr). These alloys are remarkable due to the
fact that their hydriding properties can be ‘fine‐tuned’ by alloying A or B with other
transition metal elements. The hydrides are reversible with good kinetics and moderate
operating pressures.
Hydrogen Storage
One example of a well‐studied material is the last entry in Table 11.2, v.s. titanium–
iron hydride. In terms of mass, this does not appear to be a very promising material.
Rather, it is the volumetric measure that gives such hydrides an advantage over other
storage methods. Among all the examples in Table 11.2, the titanium–iron hydride
requires one of the lowest volumes to store 1 kg. This hydride actually holds more
hydrogen per unit volume than pure liquid hydrogen and this appear counterintuitive.
In liquid hydrogen the molecules have a relatively high mobility that gives rise to the
relatively low density of the material — only 0.07 kg L−1. By contrast, hydrogen molecules
that are bonded to metal atoms in a hydride are bound closer together and thus give rise
to the high storage capacity per unit volume, despite the fact that the density of the
hydride material is higher at 5.47 kg L−1.
A typical reversible metal hydride system functions as follows. Hydrogen is supplied
at slightly above atmospheric pressure to the metal alloy (in the form of a powder) inside
a container. The reaction (11.1) proceeds to the right to form the hydride. Given that the
process is mildly exothermic, for large systems, the hydride container has to be cooled.
The hydride‐forming stage takes place at approximately constant pressure and will
require a few minutes, depending on the size of the system and how well the container
is cooled. In this context, if the pressure of gas in the container P (or, as a rule, log P) is
plotted versus the ratio of the quantity of hydrogen absorbed, Hads, to the maximum
uptake of the material, M, then an absorption isotherm is obtained and is referred to as
a ‘pressure composition isotherm (PCT)’ curve. An example of the typical behaviour
exhibited by metal hydrides is shown in Figure 11.5. Once all the metal has reacted with
the hydrogen, i.e., Hads/M = 1.0, the pressure in the container will begin to rise. At this
point, the hydrogen supply is disconnected as the vessel has reached its capacity and
needs to be sealed.
When the stored hydrogen is required, the vessel is connected to, for example, a
PEMFC. Hydrogen will be released so long as the pressure of the fuel cell is lower than
the desorption pressure, Pd. If the pressure rises above Pd, the reaction will slow down
or stop. The process is now endothermic, so heat must be provided. This is taken from
5.8
5.6
Log P (P in Pa)
Figure 11.5 PCT curve for hydrogen
absorption and desorption with a
LaNi5 alloy. Note that absorption and
desorption have different pressures
(Pa, Pd) in the almost horizontal
portion of the relationship; pressure
is given in Pa.
Pa
5.4
5.2
5.0
0.0
Pd
0.2
0.4
0.6
Hads / M
0.8
1.0
335
Fuel Cell Systems Explained
the surroundings — the vessel will cool slightly during discharge of the hydrogen. It can
be warmed slightly to increase the rate of supply by using, for example, the hot water or
air from the cooling system of the fuel cell. Once the reaction has been completed and
all the hydrogen has been withdrawn, the whole procedure can be repeated.
Different hydrides will exhibit their own characteristic PCT. Each curve usually
demonstrates hysteresis between absorption and desorption, i.e., Pa is generally higher
than Pd. Also note that the useful part of the isotherm is the almost horizontal section
that shows the range of hydrogen uptake (Hads/M) over which hydrogen can be absorbed
and desorbed. This range is always less than the maximum amount that can be absorbed
by the material.
The enthalpy change, ΔH, for the hydriding reaction (11.1) can be related to the
equilibrium dissociation pressure, P, of the hydride by the van’t Hoff 5 equation, i.e.,
ln P
H
RT
S
R
(11.2)
where ΔS is the change in entropy, T is the absolute temperature and R is the gas
constant. A logarithmic plot of the dissociation pressure against the reciprocal of the
absolute temperature should therefore be linear with a slope that is a measure of the
heat that has to be supplied to form or decompose the hydride. A series of plots for
various hydrides over the temperature range −20 to 400°C is presented in Figure 11.6.
The data clearly show the LaNi5H6, TeFeH and MmNi5H6 hydrides exhibit good
700
500
Temperature/°C
200 150
100
300
10
6
Na3AIH6
Mg2FeH6
50
2
20
0
–20
Na3AIH4
MmNi5H6
4
Pressure/MPa
336
TiCr1.8H1.7
Mg2NiH4
PdH0.6
1
CaNi5H4
TiFeH
6
4
2
0.1
6
4
LaNi5H6
MgH2
2
LaNi3.5AI1.5H5
0.01
1.0
1.5
LaNi4AIH5
2.0
2.5
3.0
1000/Temperature/K–1
3.5
4.0
Figure 11.6 Dissociation pressures of various metal hydrides. Note that Mm in MmNi5H6 denotes
mischmetal, which is a mixture of alkali earth elements.
5 Jacobus Henricus van’t Hoff (1852–1911) was one of the founders of modern physical chemistry. He was
the first recipient of the Nobel Prize in Chemistry in 1901.
Hydrogen Storage
characteristics at close to normal temperatures and pressures. These are examples
of materials that have received the most attention for application in low‐temperature
fuel cells. High‐temperature materials, such as the various magnesium hydrides, are
attractive in that the storage capacity is greater than that for low‐temperature materials.
They are, however, more prone to degradation due to the repeated expansion and
contraction of the metal lattice during adsorption–desorption cycling. Indeed, provision
must be made in the design of container to allow for this behaviour.
Much research has been carried out on improving both the long‐term performance
and the kinetics of absorption–desorption by exploring various metal alloys. For example,
alloying magnesium and nickel to give Mg2Ni provides a material that is more easily
activated than pure magnesium to yield the hydride Mg2NiH4 (3.6 wt.% H2). Of the
hydride materials that have been explored for hydrogen storage at low and medium
temperatures, the alloys of aluminium (alanates) have probably received the most attention
from researchers.
Generally, hydride materials can sustain several hundred charge–discharge cycles.
Nevertheless, as with rechargeable batteries, these systems can be abused. For example,
if the hydride is filled at too high pressure, the charging reaction will proceed too fast,
and the material may overheat and degrade. Hydrides can also be damaged by impurities
in the hydrogen, for instance, hydrogen produced by an alkaline electrolyser must be
purified to remove any traces of water, alkali or oxygen.
The vessel in which the hydride is contained should be able to withstand a reasonably
high pressure, especially if it is likely to be filled from a high‐pressure supply. For example,
the small HydrostikTM unit shown in Figure 11.7 and manufactured by Horizon fuel
cells has a rated charging pressure of 3.0 MPa. The unit is intended to be used with a
‘MiniPak’ USB charger for fuel cells, as discussed in Section 4.9.1, Chapter 4. The
Horizon technology employs an AB5‐type metal hydride with an expected lifetime of
10 years and, it is claimed, a stored energy that is equivalent to 10 disposable AA
batteries, but of course the unit can be recharged. Further details are given in Table 11.6.
The volumetric measure (i.e., mass of hydrogen per litre) is nearly as good as that
for LH2, and the gravimetric measure is very much the same as for a small cylinder of
compressed gas.
Figure 11.7 Hydrostik hydrogen canister (left) that accompanies the Horizon ‘MiniPak’ fuel‐cell USB
charger (right), employed to power portable electronic equipment such as smartphones. (Source:
Reproduced with permission of Horizon fuel cells.)
337
338
Fuel Cell Systems Explained
Table 11.6 Details of Horizon Hydrostik hydrogen container
for portable electronics equipment.
Dimensions of container (mm)
22 (diameter) × 8 (length)
Mass of empty container (g)
105
Mass of hydrogen stored (g)
1
Storage efficiency (wt.% H2)
0.9
−1
Specific energy (Wh g )
0.133
Volume of tank (approximately, L)
0.033
Mass of H2 (kg L−1)
0.03
Proponents of hydrogen storage via a reversible metal hydride note that it is safer than
a pressurized cylinder that could rapidly and dangerously discharge in the event of
failure or damage. If a leak develops in a hydride vessel, the temperature of the hydride
will fall, which will inhibit the release of the gas. It should, however, be cautioned that
there could be ingress of air that could lead to combustion of the hydride powder.
When coupled with a fuel cell, the low pressure required by hydrides helps to simplify
the design of the hydrogen supply system. Consequently, hydrides are attractive for a
very wide range of applications where small quantities of hydrogen are stored. The
materials are also particularly suited to applications where space rather than weight is a
problem. In fuel‐cell powered boats, for instance, the hydride‐storage vessel can be
located near the bottom of the hull where additional weight is often an advantage but
where space is at a premium.
The disadvantages of reversible hydrides are particularly noticeable when larger
quantities of hydrogen are to be stored, for example, in vehicles. The specific energy is
poor. Also, the problem of heating during filling and cooling during the release of
hydrogen becomes more acute as the amount of hydride increases. Consequently, the
reactant bed should have a high thermal conductivity. Large systems have been tested
for vehicles, and a typical refill time is about 1 h for a tank of approximately 5 kg. As
noted previously, the other main disadvantage of metal hydrides is that the hydrogen
must be of very high purity.
11.6
Simple Hydrogen‐Bearing Chemicals
11.6.1
Organic Chemicals
Certain organic chemicals contain significant atomic proportions of hydrogen that can be
recovered and therefore may be considered as prospective hydrogen‐storage materials.
Cyclohexane (C6H12), for example, has been proposed as it is easily decomposed
catalytically into hydrogen and benzene (C6H6) according to:
C 6 H12
C 6 H6 3H2
(11.3)
Though, in theory, this reaction can be carried out in the gas phase cleanly at a moderate
temperature, it is normally achieved over a catalyst at 500–600°C. The yield is high,
Hydrogen Storage
but there is a risk of cracking to give unwanted by‐products and the nature of the
products is dependent on the catalyst. By contrast, the reverse reaction (i.e., the
reaction between benzene and hydrogen) readily occurs at quite modest temperatures
(150–200°C) either in the liquid or vapour phase over a platinum catalyst. Both benzene
and cyclohexane are liquids at ambient temperature and pressure. The amount of
hydrogen stored by the conversion of cyclohexane to benzene would be 7.1 wt.% and
therefore has received serious attention from researchers, as has methylcyclohexane
(C7H8) and decalin (C10H18) with a yield of 6.1 and 7.2 wt.% H2, respectively.
Until recently, the high temperatures involved, control of the reactions and the
risk of by‐product formation have generally ruled out such materials for on‐board
vehicles. Nonetheless, a modified form of cyclohexane — bis‐BN cyclohexane
(C2B2N2H12) — may offer better prospects. This material, which appears to be remarkably
stable up to 150°C and has a storage capacity of 4.7 wt.% H2,6 has been found to
decompose easily at room temperature over a catalyst to yield hydrogen and no
detectable by‐products.
Heterocyclic compounds, principally n‐ethylcarbazole, and dibenzyl toluene, have
been investigated as storage materials. Although the potential storage efficiency for such
compounds is high, there are a myriad of issues in terms of their practical application,
such as low reaction kinetics, toxicity and difficulty in efficiently reversing the dehydrogenation. In fact, reversing dehydrogenation has ruled out several organic chemicals.
For example, formic acid stores 4.3 wt.% H2 and can be decomposed over a catalyst into
hydrogen and carbon monoxide (CO), but reversing the process, i.e., generating formic
acid from hydrogen and CO or CO2 is not a trivial process.
11.6.2
Alkali Metal Hydrides
A calcium hydride system for hydrogen production has been proposed.7 The reaction is
as follows:
CaH2 2H2O
Ca OH
2
2H 2
(11.4)
It could be said that the hydrogen is being released from the water by the hydride.
Both sodium and lithium hydrides also react with water to release hydrogen and were
investigated as storage materials in the late 1990s with support from the US Department
of Energy. In each case, the alkali metal hydride is highly reactive and has to be protected
from accidental contact with water from the atmosphere. Accordingly, one proposal
was to encase pellets of sodium hydride in polyethylene and then cut them open under
water. Another approach was to slurry lithium hydride with an organic material such as
a light mineral oil. The hydrogen content of this lithium hydride is the highest for any
hydride and is three times that of sodium hydride.
6 Chen, G, Zakharov, LN, Bowden, ME, Karkamkar, AJ, Whittemore, SM, Garner, EB, Mikulas, TC, Dixon,
DA, Autrey, T and Liu, S‐Y, 2015, Bis‐BN cyclohexane: a remarkably kinetically stable chemical hydrogen
storage material, Journal of the American Chemical Society, vol. 137(1), pp. 134–137.
7 Bossel, UG, 1999, Portable fuel cell battery charger with integrated hydrogen generator, Proceedings of the
European Fuel Cell Forum Portable Fuel Cells conference, 21–24 June 1999, Lucerne, pp. 79–84.
339
340
Fuel Cell Systems Explained
Unfortunately, however, no commercial products based on alkali metal hydrides have
emerged, principally on account of the difficulty and costs involved in producing the
hydride and in recycling the spent material.
11.6.3
Ammonia, Amines and Ammonia Borane
Ammonia has often been proposed as a means of hydrogen distribution and storage.
The molecular formula is NH3, which immediately indicates its potential as a hydrogen
carrier (it contains 17.7 wt.% of accessible H2). Under normal conditions, ammonia is
a highly toxic, colourless gas with a pungent choking smell that is easy to recognize. It
is produced in huge quantities — currently, at around 100 million tonnes per annum
(including a little over 16 million in the United States alone). The manufacture
of fertilizer is the most important of the many uses of ammonia in the chemical
industry and accounts for about 80% of its consumption. Ammonia is a liquid at room
temperature under a few atmospheres pressure and is therefore attractive for transport
in cylinders and pipelines in the same way as liquid petroleum gas or propane is
distributed today.
The recovery of hydrogen from ammonia involves simple dissociation, i.e.,
2NH3
N 2 3H2
H
46.4 kJmol
1
(11.5)
For this reaction to occur at a practical rate, the ammonia has to be heated to between
600 and 800°C and passed over a catalyst. Higher temperatures are necessary if the
output from the converter is to have residual levels of ammonia at the ppm level as
required by PEMFC or phosphoric acid fuel cell (PAFC) systems. Note that the
dissociation is endothermic, and if the ammonia is supplied initially as liquid, a
significant amount of energy is also required to vapourize it into a gas (+ΔH = 23.3 kJ mol−1),
which is why it is employed as a refrigerant. If the molar specific heat of ammonia is
taken to be 36.4 kJ mol−1, then the heat required to raise the temperature from 0 to
800°C for reaction (11.5) is:
H
800 36.4 29.1kJmol
1
(11.6)
The process results in the production of 1.5 mol of hydrogen for every mole of NH3,
for which the molar enthalpy of formation (HHV) is −285.84 kJ mol−1. The best possible
process efficiency for the decomposition of ammonia to hydrogen would therefore be:
285.4 1.5
23.3 29.1 46.4
285.4 1.5
0.77(or 77%)
(11.7)
This should be regarded as the upper limit to efficiency for using ammonia as a storage
medium as it does not take into account the production of ammonia from hydrogen and
nitrogen. Also, the heat required to raise ammonia to the high temperature for reaction
(11.5) and the heat of reaction itself are likely to prove challenging for low‐temperature
fuel‐cell systems.
A potential game changer for ammonia as a hydrogen provider is an alternative
‘cracking’ pathway that can be conducted at a lower temperature and, moreover, not with
a precious metal catalyst but with the bulk chemical sodium amide (NaNH2) effectively
Hydrogen Storage
acting as catalyst. Recently, researchers at Oxford, United Kingdom, have demonstrated
this method by powering a small PEMFC.8 The reactions involved are:
NaNH2
Na s
Na s
NH3 g
1
N2 g
2
NaNH s
H2 g
1
H2 g
2
(11.8)
(11.9)
(s, solid state; g, gaseous state)
Essentially, an alkaline imide (e.g., NaNH2) decomposes to yield the amide (e.g., NaNH)
and hydrogen. The ammonia is a ‘mediator’ in the two reactions. Subsequent experiments
have shown that imide‐forming amides are highly active, with the lithium amide–imide
system exhibiting superior activity per unit mass than the sodium counterpart.
Two other issues arise when decomposing ammonia. The first is that nitrogen formed by
the decomposition reaction will need to be separated from the hydrogen. Otherwise, the
nitrogen will act as a diluent in the fuel cell and cause a loss of system efficiency, as
mentioned in Section 10.5, Chapter 10. The second problem is that if a PEMFC or a PAFC is
being operated, then any ammonia remaining in the product gas can potentially react with
the acid electrolyte in these fuel cells and thereby lead to eventual failure of the system.
One method of overcoming some of the difficulties of using ammonia in liquid form
is to convert it into an ammine such as Mg(NH3)6Cl2, which is an inert solid that holds
51 wt.% NH3 and thus 9.1 wt.% H2. This compound is safe to handle and can be
compacted into dense tablets that have a low vapour pressure of ammonia at ambient
temperature and, on a volumetric basis, contain 60–70% more hydrogen than LH2.
In contrast to ammonia, hydrazine hydrate (N2H4∙H2O), which has a recoverable
hydrogen content of 80 wt.%, is an endothermic compound and therefore easier to
decompose than ammonia. Indeed, it has been known to decompose explosively.
Although difficult to manufacture or handle in bulk, hydrazine has served as a rocket
fuel and, many years ago, was tested in an experimental alkaline fuel cell.
Ammonia borane (NH3BH3) has long been promoted as a hydrogen carrier and storage
medium. The molecule of ammonia borane is similar in chemical character to that of
ethane (CH3CH3) but is a solid rather than a gas. On decomposition at 100–200°C, it
yields up to 12 wt.% H2 and thereby forms polymeric iminoborane (NHBH)n that can, in
principle, be reconverted to ammonia borane. In practice, however, the form of the
polymer is highly dependent on the decomposition conditions and the choice of catalyst
so that the reconversion process presents a major chemical challenge.9
11.7
Complex Chemical Hydrides
There is a class of inorganic metal hydrides that are ionic rather than metallic in nature.
For instance, the elements boron and aluminium form the hydride ions [BH4]− and
[AlBH4]−, respectively. When combined with alkali metal cations, soluble ionic salts are
8 Hunter, H, Makepeace, J, Wood, T, Kibble, M, Nutter, J, Jones, M and David, B, 2015, Demonstrating
hydrogen production from ammonia — powering a 100 W PEM fuel cell, Proceedings of the World
Hydrogen Technology Convention, 11–14 October 2015, Sydney.
9 Peng, B and Chen J, 2008, Ammonia borane as an efficient and lightweight hydrogen storage medium,
Energy and Environmental Science, vol. 1, pp. 479–483.
341
342
Fuel Cell Systems Explained
formed, e.g., LiBH4, LiAlH4, NaBH4 and NaAlBH4. These compounds are generally
known as ‘complex chemical hydrides’. Lithium and sodium borohydrides and the
corresponding alumino hydrides are used in organic chemistry as reducing agents.
11.7.1
Alanates
For hydrogen storage, the aluminohydrides (the so‐called alanates) are generally
preferred to the borohydrides. Thermal decomposition of the alanate NaAlH4 takes
place in two steps as follows:
3NaAlH 4
Na 3 AlH6 2Al 3H2
Na 3 AlH6
3NaH Al
3
H2
2
(11.10)
(11.11)
The reactions for the pure compounds are reversible, but only at temperatures above
the melting point of NaAlH4 (183°C) and at hydrogen pressures of 10–40 MPa, which
are impracticable. Fortunately, the temperatures for discharge and recharge of hydrogen
may be reduced significantly by the inclusion of a titanium catalyst in the alanate.
Titanium‐catalysed NaAlH4 has thermodynamic properties that are comparable with
those of the classical low‐temperature hydrides, e.g., LaNi5H6 and TiFeH (see
Figure 11.6). Reaction (11.10) is carried out at 50–100°C and corresponds to the release
of 3.7 wt.% H2 and the second step, reaction (11.11), at 130–180°C yields a further
1.9 wt.% H2. Moreover, even if only the first reaction step can be utilized, the gravimetric
hydrogen‐storage density of NaAlH4 is greater than that offered by most of the simple
metal hydrides. By contrast, since the Na3AlH6 requires higher temperatures for hydrogen
liberation, it might prove suitable for applications other than fuel cells, such as heat
pumps and heat storage. The study of NaAlH4 is still at a preliminary stage, and there is
considerable scope for the development of alternative catalysts and the optimization of
their performance. Two of the main problems to be addressed with these materials are
that they are pyrophoric (i.e., their propensity to ignite spontaneously) and they are
costly to produce.
11.7.2
Borohydrides
In recent years, developers of fuel cells have shown considerable interest in sodium
tetrahydroborate, which is more commonly referred to as sodium borohydride, NaBH4.
This compound was introduced in Section 5.2.3, Chapter 5, as a potential fuel for AFCs
and again in Section 6.5, Chapter 6, as the fuel for borohydride fuel cells. As a general
rule, the alkali metal borohydrides contain more hydrogen than the alanates, e.g., 18.5
and 10.6 wt.% H2 for the respective lithium (LiBH4) and sodium (NaBH4) analogues,
but are more stable and therefore less practical as accessible hydrogen carriers. The
compounds only decompose at relatively high temperatures (LiBH4 above 300°C and
NaBH4 above 350°C), but it may be possible to lower these temperatures by incorporating
catalysts in the materials.
Sodium borohydride can be supplied as a solid, in which case it is often mixed with
cobalt chloride, which acts as a desiccant. The material is hazardous and can spontaneously
Hydrogen Storage
give off hydrogen if it accidentally comes into contact with water. It is, however, able to
dissolve without reaction in aqueous alkaline solutions (e.g., sodium hydroxide) and
in this form is stable for long periods. Given the limitations with the decomposition of
the solid form, i.e., the high temperature requirement and use of catalyst, sodium
borohydride has attracted the most attention through its reaction with water to form
hydrogen, as expressed by:
3NaBH 4 2H2O
NaBO2 4 H2
H
218kJmol
1
(11.12)
The reaction is not reversible but has the advantage that 50% of the hydrogen comes
from the water — in effect, NaBH4 is a ‘water‐splitting’ agent. Given that alkaline
solutions of NaBH4 in water are quite stable, a catalyst is usually required to promote
the decomposition, and consequently the generation of hydrogen is quite controllable.
Notable features of reaction (11.12) are as follows:
●
●
●
It is exothermic, at a rate of 54.5 kJ mol−1 of hydrogen.
Hydrogen is the only gas produced, i.e., no gas separation is necessary as may the
case, for example, with ammonia or methanol decomposition.
If the water is heated, then water vapour will be mixed with the hydrogen, which is a
desirable feature for PEMFC systems.
Weaker solutions of sodium borohydride are more stable than strong solutions, but
their effectiveness as a hydrogen carrier diminishes. Such solutions of borohydride are
also stable for long periods, though hydrogen evolution does occur slowly. The ‘half‐life’
of such solutions has empirically been shown to follow the relationship:
log10 (t 1 2 ) pH
0.34T 192
(11.13)
where the half‐life (t½) is in minutes and the temperature T is in Kelvin. A solution of
30 wt.% NaBH4 + 3 wt.% NaOH has a half‐life of about 2 years at 20°C. One litre of such
a solution will yield 67 g of hydrogen on hydrolysis, which is a better yield than that
obtained from any of the practical metal hydrides.
Another approach to preparing a borohydride medium for hydrogen storage is to mix
dry solid NaBH4 powder with light mineral oil and a dispersant to produce an ‘organic
slurry’. The oil coats the solid particles and protects them from inadvertent contact with
water during handling and transport and also moderates the decomposition rate of the
hydride when water is introduced. More development work is required to control the
reaction kinetics so as to yield hydrogen at the desired rate.
In addition to the previous list of advantages afforded by reaction (11.12), the following
benefits of the NaBH4 solution should be recognized:
●
●
●
●
It is arguably the safest of all hydrogen‐containing liquids to transport.
Apart from cryogenic hydrogen, it is the only liquid that gives pure hydrogen as the
product.
The reactor for the release of hydrogen requires no energy input and can operate at
ambient temperature and pressure.
The rate of hydrogen production can be simply controlled.
Unfortunately, however, there are two significant disadvantages that arise in employing
borohydrides for hydrogen storage. The first is that the product — sodium borate — cannot
343
344
Fuel Cell Systems Explained
be reused in situ, which means that it has to be replaced with fresh material once all of
the hydrogen has been released. It has been shown that this procedure on a vehicle
could be quite rapid with the borate transported to a processing plant for regeneration.
Nevertheless, the regeneration process is not only expensive but also requires a large
amount of energy and, therefore, constitutes the major obstacle to the widespread
uptake of borohydrides. It would appear that the cost of the regeneration has to fall by
a factor of approximately 50 for NaBH4 to be acceptable as a fuel for vehicles. Nonetheless,
from the standpoint of both mass and volume, the system appears superficially satisfactory
as a hydrogen‐storage scheme for FCVs. DaimlerChrysler demonstrated that a NaBH4
system, designed by Millennium Cell in the United States, could provide a minivan (the
Natrium) with a range of 480 km. Despite this performance, however, Millennium Cell
ceased development of the NaBH4 system in 2008.
In the past few years, the challenge of using borohydrides has been taken up by Cella
Energy Ltd in the United Kingdom. The hydrogen‐storage material developed by this
company is based on a proprietary complex chemical hydride that is formed into
pellets with a polymer admixture that provide mechanical integrity to the material.
The pellets are coated with a thin (<50 µm) film of a common hydrogen‐permeable or
permselective polymer, e.g., polyethylene or polymethyl methacrylate. The protective
film allows only hydrogen to diffuse out of the material once it is heated. The Cella
Energy material can be easily transported and dispensed into vehicle fuel tanks. The
protective polymer film also prevents ingress of oxygen or water and thereby gives the
material a long shelf-life.
11.8
Nanostructured Materials
In 1998, a study was published on the absorption of hydrogen in carbon nanofibres.10
The authors presented results that suggested that these materials could absorb in excess
of 67 wt.% H2. This amazingly large amount took the academic world by storm, as it
offered the prospects of levels of hydrogen storage in material at ambient temperatures
and pressures that would ensure hydrogen a certain future for vehicles. The initial
euphoria was tempered when other research groups tried to repeat the findings and
found that there were measurement errors due to the presence of metal contaminants
and/or water absorption. Nevertheless, so‐called ‘nanostructured’ materials continue to
be evaluated for hydrogen storage and other catalytic properties.
Nanotechnology has become a broad area of applied science that focuses on the
control and exploitation of materials with characteristic geometric dimensions below
10−7 m and new properties that result from the nanostructure. It draws on fields as
diverse as colloid science, device physics, molecular biology and supramolecular
chemistry to address a wide range of potential applications. By virtue of their large
surface‐to‐volume ratios, certain nanomaterials can adsorb considerable amounts of
hydrogen in the molecular state via weak molecule–surface interactions (so‐called
physisorption). This is in contrast to the chemisorption process on metal and complex
10 Chambers, A, Park, C, Baker, RTK and Rodriguez, NM, 1998, Hydrogen storage in graphite nanofibre,
Journal of Physical Chemistry B, vol. 102, pp. 4253–4256.
Hydrogen Storage
hydrides in which the hydrogen is dissociated into atoms that chemically bond with the
lattice of the storage medium. Obviously, physisorption is preferred as it would moderate the temperature and pressure required for the respective uptake and release of
hydrogen. Moreover, there is no major heat‐transfer problem because physisorption
bonds are weak, typically with enthalpies of adsorption from −10 to −20 kJ mol−1 H2. On
the other hand, significant uptake of hydrogen is normally seen only at cryogenic temperatures, and this is a major inconvenience.
Particular attention has been paid to the possibility of hydrogen storage in carbon‐
based materials that take the form of nanofibres or nanotubes. These are structures
that derive from the fundamental carbon entity C60 (buckminsterfullerene) that has a
spherical cage‐like structure that is made up of hexagons and pentagons, as shown
schematically in Figure 11.8a.11 The generic term ‘fullerene’ is used to describe a pure
carbon molecule that consists of an empty cage of 60 or more carbon atoms. Graphitic
nanofibres are prepared by the decomposition of hydrocarbons or carbon monoxide
over metal catalyst and are made up of graphene sheets aligned in a particular direction
as determined by the choice of catalyst. Three distinct structures may be formed:
platelet, ribbon and herringbone (see Figure 11.8b). The structures are flexible and can
expand to accommodate the hydrogen. Graphitic nanofibres vary from around 50 to
1000 nm in length and 5 to 100 nm in diameter.
Carbon nanotubes are cylindrical or toroidal varieties of fullerene and have lengths of
between 10 and 100 µm. ‘Single‐walled’ nanotubes are composed of only one graphene
sheet and have typical diameters of up to 2 nm. ‘Multiwalled’ nanotubes consist of
concentric rings (diameters of 30–50 nm) or spirals of graphene sheets; different types
of nanotube are illustrated in Figure 11.8c. Carbon nanotubes were first identified by
Iijkima in 1991 who obtained them by accident when using an electric arc drawn
between two carbon electrodes. Nowadays, laser ablation and chemical vapour deposition,
which are more controlled procedures, are employed. Unfortunately, such high‐tech
methods of preparation mean that carbon nanotubes are expensive materials.
Since the first edition of this textbook in 2001, many groups have investigated carbon
nanofibres and nanotubes with mixed results. One of the problems is the difficulty in
determining hydrogen uptake experimentally on small amounts of materials, largely
due to inaccuracies in measuring sample volumes.12 Whereas physisorption has been
clearly demonstrated, useful levels of hydrogen uptake (up to about 3 wt.%) have only
been achieved at cryogenic temperatures with materials of very high surface area. Some
workers claim that uptakes of about 1 wt.% H2 can be achieved on materials doped with
titanium or platinum, but long‐term consistent performance has yet to be achieved for
carbon nanofibres and nanotubes. Molecular modelling has suggested that graphene
sheets, perhaps spaced wider than in graphite and suitably functionalized, could provide
the required characteristics to meet the targets for hydrogen storage at ambient
conditions, but this prediction has not been verified experimentally.13
11 In theory, a molecule of buckminsterfullerene can be hydrogenated up to C60H60, which corresponds to an
absorption of almost 7.7 wt.% H2. In the United States, MER Corporation has demonstrated absorption up to
6.7 wt.% H2 on fullerenes, but the absorption is very slow and to date no practical application has been found.
12 Webb, CJ and Gray, EMacA, 2014, The effect of inaccurate volume calibrations on hydrogen uptake
measured by the Sieverts method, International Journal of Hydrogen Energy, vol. 39, pp. 2168–2174.
13 Tozzini, V and Pellegrini, V, 2013, Prospects for hydrogen storage in graphene, Physical Chemistry
Chemical Physics, vol. 15, pp. 80–89.
345
346
Fuel Cell Systems Explained
(a)
(c)
(b)
Platelet
Ribbon
Herringbone
(d)
Figure 11.8 Schematic representations of (a) buckminsterfullerene, (b) carbon nanofibres, (c)
single‐ and multiwalled carbon nanotubes and (d) metal oxide framework.
In parallel with research on carbon, many other porous materials and composites
with high surface areas are being investigated as possible storage media. These include
the following:
●
●
Zeolites: Complex aluminosilicates with engineered pore sizes and high surface areas.
Metal–organic frameworks (MOFs): Typically zinc oxide structures bridged with
benzene rings; an example is illustrated schematically in Figure 11.8d. These
materials have extremely high surface areas, are highly versatile and allow for many
Hydrogen Storage
●
structural modifications. As with carbon structures, MOFs tend to absorb significant
quantities of hydrogen only at high pressure and cryogenic temperatures, typically
77 K.
Clathrate hydrates: Water (ice) cage‐like structures that often contain guest molecules
such as methane and carbon dioxide (see Section 10.2.2, Chapter 10). Hydrogen caged
in a clathrate hydrate was first reported in 2002 and requires very high pressures to be
stable. In 2004, researchers showed solid H2‐containing hydrates could be formed at
ambient temperature and high pressure by adding small amounts of promoting
substances such as tetrahydrofuran.14 These clathrates have a theoretical maximum
hydrogen absorption of around 5 wt.% and 40 kg m−3.
11.9
Evaluation of Hydrogen Storage Methods
Fuel‐cell vehicles are the most demanding application for hydrogen storage. At
present, the only practical means of storing hydrogen on a road vehicle is as a high‐
pressure gas in a cylinder. As discussed previously, modern lightweight composite
cylinders are a distinct improvement on conventional steel vessels. Other options
such as cryogenic liquid hydrogen, organic chemicals or metal hydrides are likely to
be uneconomic, at least in the short-term. Irreversible chemical hydrides that react
with water may find some success, but manufacturing and recycle costs need to be
reduced dramatically. To this end, further investigation is necessary both at the
fundamental scientific level and in terms of vehicle engineering, hydrogen recovery
plant and overall logistics.
The task of displacing liquid hydrocarbon fuels by hydrogen is formidable. This is
illustrated pictorially in Figure 11.9 and numerically in Table 11.7. The latter presents
a comparison of the probable masses and volumes of systems to accommodate the
amount of hydrogen (11.75 kg) that is equivalent in energy content (1.4 GJ) to 45 L of
gasoline, which is typically the capacity of the fuel tank in a car. Columns 3 and 5
relate these masses and volumes, respectively, with gasoline taken as unity. In the
case of methanol, the mass and volume of an on‐board reformer is also taken into
account. The data may only be approximate, because of the complexity of making due
allowance for the containers and ancillary equipment, but the information does give
some indication of the magnitude of the problem. It is seen that gasoline is well ahead
of the various options for hydrogen storage, both in terms of mass and volume. The
analysis also reveals why on‐board methanol reforming has been seriously considered
for FCVs and why borohydride solutions would be attractive were it not for the
difficulty in handling the material and the cost of regeneration. In this context, it
should be pointed out that the tank‐to‐wheel efficiency (see Section 12.4.1,
Chapter 12) of a fuel‐cell car is significantly better than that of either a conventional
gasoline or diesel engine.15
14 Florusse, LJ, Peters, CJ, Schoonman, J, Hester, KC, Koh, CA, Dec, SF, Marsh, KN and Sloan, ED, 2004,
Stable low‐pressure hydrogen clusters stored in a binary clathrate hydrate, Science, vol. 306(5695), pp. 469–471.
15 Davis, C, Edelstein, W, Evenson, W, Brecher, A and Cox, D, 2003, Hydrogen Fuel Cell Vehicle Study,
A Report Prepared for the Panel on Public Affairs (POPA), American Physical Society, College Park, MD.
347
348
Fuel Cell Systems Explained
Figure 11.9 Schematic representation of the relative volumes of two hydrides, liquid hydrogen and
compressed gas (20 MPa pressure) to contain 4 kg of hydrogen for a vehicle driving range of 400 km.
Table 11.7 Approximate comparison of mass and volume of gasoline and stored forms
of hydrogen for equivalent energy content (1.4 GJ).
Mass of store
and fuel (kg)
Gasoline (45 L)
Index
Volume of
store (L)
Index
41
1
45
1
Compressed H2 (20 MPa)
conventional steel cylinder
~1150
28
~1080
24
Compressed H2 (70 MPa)
composite cylinder
~200
4.9
~170
3.8
LH2 in cryostat
~100
2.4
~350
7.8
Ti–Fe hydride bed
~1050
25.6
~275
~6.1
Methanol and reformer
~1140
27.7
~294
~6.5
NaBH4 solution
~553
13.5
192
~4.3
In conclusion, mention should be made of energy stored in rechargeable batteries
and how this compares with energy stored as hydrogen. As an example, consider the
small metal hydride container shown in Figure 11.10, which holds 1.7 g of hydrogen.
Taking the data given in Section A2.4, Appendix 2, for a PEMFC operating at
Hydrogen Storage
Figure 11.10 Small metal hydride hydrogen store for fuel cells used with small portable electronic
equipment.
Vc = 0.6 V, which corresponds to an efficiency of 40% (HHV), the hydride container
will deliver 26.8 0.6 1.7 27 Wh of electrical energy (the effective specific electrical
energy of hydrogen given in Table A2.1 is 26.8 × Vc kWh kg−1). This is approximately
the same as the capacity of six D‐size nickel–cadmium cells, each with roughly
the same volume as the hydrogen store of Figure 11.10. In other words, the stored
hydrogen has obtained approximately six times the energy density (Wh L−1) of
nickel–cadmium batteries.
The most advanced lithium‐ion batteries exhibit a specific energy of around
900 kJ kg−1. Hydrogen itself has a specific energy of 142 MJ kg−1, which would indicate
that hydrogen offers an amazingly better energy storage option. When the weight of
the storage vessel is taken into account the figures are less compelling. Manipulation
of the data in Table 11.7 will show that compressed hydrogen in a conventional
cylinder at 20 MPa has a specific energy of some 1.2 MJ kg−1. If the storage pressure is
raised to 70 MPa, the specific energy is increased to 7 MJ kg−1. This is still eight times
the specific energy of the lithium‐ion battery and is one of the reasons why hydrogen is
such an attractive means of storing electrical energy on vehicles. Further consideration
shows that the specific energy of hydrogen stored in the form of hydrides is
potentially even higher, but, as mentioned earlier, there are other considerations
(principally heat management and slow reaction rates) that make hydrides currently
a poor choice for storage on vehicles.
Fuel processing and hydrogen storage are vital aspects of fuel‐cell system design,
but they are not the only subsystems that have to be added to the fuel‐cell stack. The
next chapter addresses issues associated with moving the reactant gases through
the system.
349
350
Fuel Cell Systems Explained
Further Reading
Broom, DP, 2011, Hydrogen Storage Materials — The Characterisation of Their Storage
Properties (Green Energy and Technology), Springer‐Verlag, London. ISBN:
978‐0‐85729‐220‐9.
Gray, EMacA, 2007, Hydrogen storage — status and prospects, Advances in Applied
Ceramics, vol. 106(1–2), pp. 25–28.
Hirscher, M, ed., 2010, Handbook of Hydrogen Storage: New Materials for Future Energy
Storage, Wiley‐VCH Verlag GmbH, Weinheim. ISBN: 978‐3‐527‐62981‐7.
Rand, DAJ and Dell, RM, 2008, Hydrogen Energy — Challenges and Prospects, RSC
Publishing, Cambridge. ISBN: 978‐0‐85404‐597‐6.
Thomas, CE, 2009, Fuel cell and battery electric vehicles compared, International Journal
of Hydrogen Energy, vol. 34, pp. 6005–6020.
Varin, RA, Czujko, T and Wronski, ZS, 2009, Nanomaterials for Solid State Hydrogen
Storage (Fuel Cells and Hydrogen Energy), Springer, New York. ISBN:
978‐0‐387‐77711‐5.
Walker, G, ed., 2008, Solid‐State Hydrogen Storage: Materials and Chemistry, Woodhead
Publishing, Cambridge. ebook ISBN: 9781845694944.
Zhang, JZ, Li, J, Li, Y and Zhao, Y, 2014, Hydrogen Generation, Storage and Utilization,
Wiley‐Science Wise Co‐publication, John Wiley & Sons, Inc., New York. ISBN:
978‐1‐118‐14063‐5.
351
12
The Complete System and Its Future
It has been said that a fuel‐cell system comprises three elements: (i) the processor for
providing the fuel to the stack, (ii) the stack itself and (iii) the power conditioner for
converting raw DC power from the stack to a useful AC voltage. The stack is the heart
of the system, but many other components are required for it to function in a real‐world
application. For instance, balance‐of‐plant (BoP) items are necessary to provide cooling
by moving air around the system and to supply oxygen to the cathode. Pumps, fan
compressors and blowers are required to deliver this service. In addition, the energy in
the stack exhaust gases can sometimes be harnessed instead of simply going to waste. The
technology for such mechanical equipment is very mature given its development and
long‐term use for other applications. Designers of fuel‐cell systems will therefore choose
devices principally custom‐made for products in other markets. ‘Gas‐moving devices’
can vary greatly in size and application; therefore, in considering their suitability for a
given fuel‐cell system, it is necessary to examine a wide range of candidate equipment.
Consequently, the first section of this chapter will consider the various mechanical BoP
items required to move both air and fuel gas, i.e., compressors, turbines, ejectors, fans/
blowers and pumps. Section 12.2.2 then addresses issues associated with electrical
components, or how the DC power from a fuel‐cell stack is converted to a more useful
AC. Section 12.4.1 covers the integration of batteries with fuel cells — an area of growing
importance not only for road vehicles but also for stationary renewable power systems.
Section 12.4.2 presents an analysis of fuel‐cell systems and brings together information
from all of the previous 11 chapters. The final sections review the commercial status
and future prospects of complete fuel‐cell systems.
12.1
Mechanical Balance‐of‐Plant Components
12.1.1
Compressors
The principal operation of each of the four main types of compressor that are used in
fuel‐cell systems is represented schematically in Figure 12.1. The basic Roots compressor,
depicted in Figure 12.1a, is one of the simplest of positive‐displacement pumps and is
frequently employed as a supercharger in diesel engines where it is driven directly
from the engine crankshaft via a belt, chain or gears. The device operates by pumping
a gas with a pair of meshing lobes, not unlike a set of stretched gears. Fluid is trapped
Fuel Cell Systems Explained, Third Edition. Andrew L. Dicks and David A. J. Rand.
© 2018 John Wiley & Sons Ltd. Published 2018 by John Wiley & Sons Ltd.
352
Fuel Cell Systems Explained
(a)
(b)
(c)
(d)
Volute
Rotor
Figure 12.1 Some different types of compressor: (a) Roots positive‐displacement pump, (b) twin‐screw
Lysholm, (c) centrifugal (radial) design and (d) axial flow design.
in pockets surrounding the lobes and is carried from the intake side to the exhaust.
The Roots compressor is quite cheap to produce and works over a wide range of flow
rates. However, it only has a modest efficiency when delivering a small pressure
lift — typically, about 90% at a pressure ratio (Pout/Pin) of 1.1 — and is therefore
normally employed for moving large volumes of gas. The performance can be enhanced
by increasing the number of rotors from 2 to 3 or 4, and by twisting the rotors by 60%
to form a partial helix. Such improvements, while increasing the cost, also greatly
reduce pressure fluctuations in the gas and provide a marginal increase in efficiency.
Even so, such compressors are still only suitable for small pressure increases, typically
up to a pressure ratio of 1.8.
The Lysholm screw compressor of Figure 12.1b has two screws, which counter rotate
and thereby drive the gas up through the region between the two screws while compressing it at the same time. The device can be thought of as a refinement of the
‘Archimedes screw’, which has been used for pumping water since ancient times. There
are two variations of the twin‐screw compressor. In the first, an external motor drives
only one rotor, and the second rotor is turned by the first. The arrangement requires
the rotors to be in contact and therefore lubricated with oil. The small amounts of oil
that is inevitably carried forward with the air are of no concern for many industrial
The Complete System and Its Future
(a)
(b)
Figure 12.2 (a) Typical centrifugal air‐compressor rotor and (b) twin‐screw compressor.
applications; the twin‐screw compressor finds wide application in providing compressed
air for pneumatic tools. In the second type of twin‐screw compressor, the two rotors
are connected by a synchronizing gear — a separate pair of cogs provides the driving
link from one rotor to the other. The counter rotating screws do not come into contact,
though for a good efficiency they will run very close to each other. This version gives
an oil‐free output, which is necessary for a fuel‐cell system; it is also used in other
facilities, e.g., for circulating the fluid in refrigeration systems.
By changing the length and pitch of the screws, the Lysholm compressor can be
designed to cover an extensive number of compression ratios — the exit pressure can be
up to eight times the input pressure. Another advantage is that the efficiency remains
high over a broad range of flow rates. Unfortunately, however, this type of compressor
is expensive to manufacture on account of the precision required by the rotors,
synchronizing gears and bearings (there are high lateral and axial mechanical loadings
on the axles of screw compressors).
One of the more common types of compressor or blower is the centrifugal or radial
design illustrated in Figure 12.1c. The gas is drawn in at the centre of the impeller
and forced out at high speed to the surrounding volute. Here the kinetic energy is
‘converted’ into a pressure increase. The centrifugal compressor is used in the vast
majority of engine turbocharging systems. An example rotor is shown in Figure 12.2a.
Although the shape may be complex, it can usually be cast as one piece. This type of
compressor is thus low cost, well developed and available to suit a wide range of flow
rates. In addition, the efficiency compares well with that of other types of compressor,
but operation must kept within well‐defined flow rates and pressure changes to obtain
high efficiencies. Indeed, the centrifugal compressor cannot operate at all at low flow
rates, as is explained in Section 12.2. Another problem is that the rotor must rotate at
a very high speed (80 000 rpm is typical), whereas twin‐screw and Roots compressors
(such as shown in Figure 12.2b) are limited to about 15 000 rpm. Care must therefore
be taken with the design and lubrication of bearings in a centrifugal compressor.
353
354
Fuel Cell Systems Explained
P1 T1
P1 T1
P2 T2
P2 T2
Figure 12.3 Symbols for (a) a compressor and (b) a turbine.
The axial flow compressor, as shown in Figure 12.1d, drives the gas by rotating a
large number of blades at high speed. It is, in essence, the inverse of the turbine that is
commonly operated in thermal power plants. As with turbines, the gap between the
ends of the blades and the housing must be as small as possible. This requirement
adds considerably to the manufacturing cost. The efficiency is high, but only over a
fairly narrow range of flow rates. Experience with diesel systems suggests that the
axial flow compressor will be the best option for fuel cells that have an output above a
few megawatts and that are only operated between full power and half power at any
given time.
The rotating vane compressor is used for air management in some industrial
operations. The device claims advantages in terms of cost over the screw compressor.
Nevertheless, it is unlikely to be used for fuel‐cell systems, since the tips of the rotating
vanes must be lubricated with a thin film of oil, and so, even after filtering, there will
always be some oil in the output gas that, as noted earlier, is not generally acceptable.
The symbol adopted in process flow diagrams (PFDs) for compressors of all types is
as shown in Figure 12.3a with the complementary symbol for turbines shown in
Figure 12.3b.
12.1.1.1
Efficiency
As with the efficiency of a fuel cell, care needs to be taken in defining the efficiency
of a compressor. Whenever a gas is compressed, work is done on the gas and so its
temperature will rise, unless the compression is enacted very slowly or there is substantial
cooling. In a process that is reversible and also adiabatic (no heat loss), it can readily be
shown that if the pressure changes from P1 to P2, then the temperature will change from
T1 to T2′ according to the following relationship:
1
T2
T1
P2
P1
(12.1)
The Complete System and Its Future
where γ = CP/CV, i.e., the ratio of the specific heat capacities of the gas at constant
pressure (CP) and constant volume (CV). This formula gives the temperature of the gas
as a consequence of the change in pressure, as is indicated by the prime (ʹ) on T2′, for an
isentropic process, i.e., one in which the entropy of the system remains unchanged. In
practice, the new temperature, T2′, will be higher than given by equation (12.1) since
some of the motion of the moving blades and vanes will be ineffective in compressing
the gas and may be observed as a small increase in gas temperature. Also, some of the
gas might ‘churn’ around the compressor and, thereby, become hotter without being
compressed. If the actual new temperature is T2, then the ratio between the following
two quantities will enable derivation of the compressor efficiency:
1) The actual work done to raise the pressure from P1 to P2.
2) The work that would have been done if the process had been reversible or
isentropic — the ‘isentropic work’.
To find these two quantities, the following assumptions, which are generally valid, can
be made:
●
●
●
The heat flow from the compressor is negligible.
There is no change in kinetic energy of the gas as it flows into and out of the compressor,
or at least any change is negligible.
The gas is a ‘perfect gas’ and so the specific heat at constant pressure, CP, is constant.
With these assumptions, the work done, W, is simply the change in enthalpy of the gas,
namely:
W
(12.2)
C P T2 T1 m
where m is the mass of gas compressed. The isentropic work done, W′, is given by:
W
(12.3)
C P T2 T1 m
The isentropic efficiency, ηc, is the ratio of these two quantities, W′/W. Hence:
isentropic work
c
C P T2 T1 m
C P T2 T1 m
real work
T2 T1
T2 T1
(12.4)
If the isentropic temperature T2′ from equation (12.4) is substituted into equation (12.1),
then:
c
T1
(T2 T1 )
1
P2
P1
(12.5)
1
Equation (12.5) can also be rearranged to give the change in temperature on compression,
as follows:
T
T2 T1
T1
c
1
P2
P1
1
(12.6)
355
356
Fuel Cell Systems Explained
This definition of isentropic efficiency does not consider the work done on the shaft
driving the compressor. To include this work, the mechanical efficiency, ηm, should also
be considered, as this takes into account the friction in the bearings, or between the
rotors and the outer casing (if any). In the case of centrifugal and axial compressors, the
mechanical efficiency is very high, typically over 98%, so that the total efficiency, ηT, can
reasonably be expressed by:
T
m
(12.7)
c
The isentropic efficiency, ηc, is the most useful measure of efficiency because it is
related directly to the rise in temperature (ΔT), and this can be quite high. For example,
using air at 20°C (293 K) for which γ = 1.4, a doubling of the pressure and a typical value
for ηc of 0.6, when substituted into equation (12.6), gives:
T
293 0.286
2
1
0.6
170 K
(12.8)
For some fuel cells, the temperature rise is beneficial as it preheats the reactants. On
the other hand, for low‐temperature fuel cells, it means that the compressed gas needs
cooling. Such coolers, which are located between the compressor and the fuel cell, are
often called ‘intercoolers’.
12.1.1.2
Power
The power needed to drive a compressor can be readily found from the change in
temperature and knowledge of the heat capacity of the gas. Thus:
Power C P
T m
(12.9)
where ṁ is the rate of flow of gas in kg s−1. The temperature difference, ΔT, is given by
equation (12.6). Therefore
1
Power C P
T1
P2
P1
c
1
m
(12.10)
In the case of an air compressor, which is a feature of most fuel‐cell systems, the CP
for air can be taken to be 1004 J kg−1 K−1 and γ = 1.4, so that power required by the air
compressor is given by:
Power 1004
T1
c
12.1.1.3
P2
P1
0.286
1
m
(12.11)
Performance Charts
The efficiency and performance of a compressor will depend on many factors that
include the following:
●
●
Inlet pressure, P1.
Outlet pressure, P2.
The Complete System and Its Future
●
●
●
●
●
Inlet temperature, T1.
Gas density, ρ.
Gas viscosity, μ.
As flow rate, ṁ.
Compressor rotor speed, N.
With all these variables, to tabulate or draw some kind of chart of compressor performance
would clearly be a difficult, if not impossible, task. Thus, it is necessary to eliminate or
group together the variables. This exercise is often performed in the following way:
●
●
●
The inlet and outlet pressures are combined into one variable, i.e., the pressure ratio
P2/P1.
For any gas, the density is given by ρ = P/RT and can therefore be ignored as P and T
are being considered.
The viscosity of the gas, bearing in mind the limited range of gases normally used, can
also be ignored.
Further simplification is done by a process of dimensional analysis, which can be
found in textbooks on turbines and turbochargers. The result is to group together
variables in ‘non‐dimensional’ groups. The two groups are:
m
T1
P1
N
T1
and
They are known, respectively, as the ‘mass flow factor’ (MFF) and the ‘rotational speed
factor’ (RSF). They are sometimes also referred to as the ‘non‐dimensional mass flow’
and ‘rotational speed’. Charts of the efficiency for different pressure ratios and MFFs and
lines of constant RSF are plotted. The chart for a typical twin‐screw compressor
(Lysholm) is given in Figure 12.4. The lines of constant efficiency are similar to the
contours of a geographical map — instead of indicating hills, however, the lines indicate
areas of increasing efficiency of operation.
The unit generally used for P1 is the bar and for temperature, the Kelvin. The MFF can
be related to the power of a fuel cell as follows. Assuming typical operating conditions for
the fuel cell (i.e., air stoichiometry = 2, average voltage = 0.6 V), then, from equation (A2.9)
in Appendix 2, the flow rate of the air (air usage) for a 250‐kW fuel cell is:
3.58 10 7
Vc
Pe
kg s
(A2.9)
1
where Pe is the power of the fuel cell (watts), λ is the air stoichiometry and Vc is the cell
voltage. Thus:
m
3.58 10
7
2 250000
0.6
0.3 kg s
1
(12.12)
If standard conditions are assumed for the air (i.e., P1 = 1 bar, T = 298 K), the MFF is:
MFF
0.3
298
1.0
5.18 kg s 1 K1/2 bar
1
(12.13)
357
Fuel Cell Systems Explained
Rotor speed factor/rev. min–1 K–1/2
600
η=
η=0
.4
4.0
0 .5
400
Pressure ratio
358
800
η=
0.6
1000
η=
0.7
1200
.75
η=0
3.0
2.0
1.0
0
1.0
2.0
3.0
4.0
5.0
Mass flow factor/kg s–1 K1/2 bar–1
Figure 12.4 Performance chart for a typical twin‐screw compressor. A mass flow factor of 5
corresponds to the air needs of a fuel cell with an output of about 250 kW.
Consequently, the horizontal x‐axis of Figure 12.4 corresponds to the air flow
requirements of fuel cells of power approximately 0–250 kW. Similarly, if the rotor
speed factor is 1000, this will correspond to a speed of about 17 300 rpm. The use of
these ‘non‐dimensional’ quantities is standard practice in textbooks on compressors
and turbines, but do not feature in many manufacturers’ data sheets. In the latter case,
standard condition values are applied (P1 = 1.0 bar, T = 298 K), and the MFF is replaced
with the mass flow rate or even the volume flow rate, and the RSF is replaced with speed
in rev min−1 (rpm). Generally speaking, such charts will give satisfactory results — except
in the case of multistage compressors. When gas has been compressed through the first
of a series of stages, its temperature and pressure would obviously have changed markedly,
and so the MFF will be quite different, even though the actual mass flow is unchanged.
The performance chart for a typical centrifugal compressor is shown in Figure 12.5.
The chart is different in form from that of a screw‐type or Lysholm compressor. Two
points to note are as follows:
1) There are regions of high efficiency, but these are very narrow. The constant
efficiency ‘contours’ are very close together when moving across the chart for a given
pressure ratio.
2) There is a distinct ‘surge line’ and to the left of this the compressor is unstable and should
not be used. As mentioned earlier, the centrifugal compressor works by accelerating the
gas out from the centre of the unit. If there is no gas to pump (i.e., the inlet pressure is
too low), then the pressurized gas will flow back to the inlet, only to be pumped again.
This action will cause the pressure to become erratic and the gas will start to heat up.
A centrifugal compressor, therefore, must not be operated in this ‘low flow rate’ region.
If the flow rate has to be low, the pressure ratio must also be reduced.
The Complete System and Its Future
3.4
6000
K –1/2
act
or/
rev
.m
in –1
η = 0.68
5500
5000
η=
0.7
0
1
lin
e
0.7
η=
rge
2.2
Su
4500
spe
ed
f
Pressure ratio
2.6
η = 0.65
η = 0.60
η = 0.55
3.0
1.8
Ro
tor
4000
1.4
3500
3000
2500
1.0
0
1.0
2.0
3.0
Mass flow factor/kg
s–1
4.0
K1/2
bar–1
5.0
~250 kW
Figure 12.5 Performance chart for a typical centrifugal compressor.
As a consequence of these two features, it is difficult to maintain constant pressure
with centrifugal compressors without compromising efficiency. To achieve optimum
performance, the pressure should be allowed to rise and fall as the gas flow rate increases
or decreases so as to follow the maximum efficiency regions shown in Figure 12.5.
For most applications of centrifugal compressors (e.g., as turbocharger for an internal
combustion engine), the issue of variable pressure is not a problem. This is also the case
with some proton‐exchange membrane fuel cells (PEMFCs). In many fuel‐cell systems
(especially the molten carbonate fuel cell (MCFC)), however, the pressure of oxidant
and fuel gas within the stack needs to be closely matched and therefore makes the
centrifugal compressor a poor choice.
12.1.1.4
Selection
For pressure ratios in the range of 1.4 to around 3, the best option is to employ
compressors that are designed for internal combustion engines. For example, the
Eaton supercharger shown in Figure 12.6 is produced for large petrol engines with gas
flows in the range of 50–100 L s−1. The rate corresponds to a power range of 50–150 kW
for the fuel cell.1
1 The power was calculated using the equation (A2.4) derived in Appendix 2 with typical values of
stoichiometry (λ = 2) and an average cell voltage of 0.6 V.
359
360
Fuel Cell Systems Explained
Figure 12.6 Eaton supercharger. The unit is about 25 cm long and will boost pressure to between 36
and 70 kPa. (Source: Reproduced by kind permission of Eaton Corporation (http://www.
enginetechnologyinternational.com/eaton.php).)
Pulley
Figure 12.7 Lysholm compressor (Model 1200 AX). Air enters a hole (not visible) on the left and exits
via the six holes on the visible face. The pulley on the right drives the screws.
For higher pressure ratios, a twin‐screw compressor is the first choice for efficiency
and flexibility, as is demonstrated by its performance chart given in Figure 12.4. An
example of a small Lysholm twin‐screw compressor (Model 1200 AX) is shown in
Figure 12.7. This particular unit measures 260 × 176 × 120 mm, weighs just 5 kg and is
designed for flow rates up to about 0.12 kg s−1, which corresponds to about 100 kW for
a typical fuel cell operating at λ = 2 and an average cell voltage of 0.6 V. For systems of
higher power, the fuel‐cell designer can choose from a wide range of commercially
available twin‐screw compressors that have been tried and tested over the years. They
are produced in large numbers at low cost by several manufacturers for the automotive
industry, both for original equipment manufacturers (OEMs) and as after‐market
products. Twin‐screw compressors are also frequently used to replace piston
The Complete System and Its Future
compressors where large volumes of high‐pressure air are needed, either for large
industrial applications or to operate high‐power air tools such as jackhammers.
It should be noted that most automotive pumps or superchargers are driven by a
pulley from the engine crankshaft. In fuel‐cell systems, the compressor will be driven by
an electric motor. By controlling the speed of the motor, the gas flow rate and pressure
lift can be varied and thus provide another method of optimizing the stack performance
according to the load or electrical demand on the system.
12.1.2 Turbines
In a bottoming cycle for a solid oxide fuel cell (SOFC) or an MCFC system, a turbine is
used to harness the energy in the hot exhaust gas to produce additional power. In some
cases, the turbine can also turn a compressor to compress the incoming air or fuel
gas. Two types of turbines can be combined with fuel‐cell systems. The first is the
centripetal or radial turbine, which is essentially the inverse of the centrifugal
compressor discussed earlier. This technology is the preferred choice unless the power
involved is greater than about 500 kW, when the axial turbine, which is the standard
technology employed in gas and steam turbine power generation sets, may be considered.
In both cases, the symbol for the turbine is the inverse of that for the compressor and is
shown in Figure 12.3b.
It is possible to mount a turbine and compressor side by side on the same shaft. With
the surrounding housing common to both, this makes for a very compact and simple unit.
Such an arrangement is generally known as a ‘turbocharger’, because its main application
is in the supercharging of engines using a turbine driven by the exhaust gases.
The efficiency of the turbine is treated in a similar manner to that for compressors,
with the same assumptions. If the turbine works isentropically, then the outlet temperature
will fall from T1 to T2′, whereas with the compressor:
1
T2
T1
P2
P1
(12.1)
In practice, however, because some of the energy will not be transferred to the turbine
shaft, but will stay with the gas, the outlet temperature will be higher than T2′. The actual
work done will therefore be less than the isentropic work (note: for the compressor it
was more). Consequently, the isentropic efficiency of a turbine is defined as:
actual work done
c
(12.14)
isentropic work
Making the same assumptions about ideal gases, as were made for compressors,
equation (12.14) becomes:
c
T1 T2
T1 T2
T1 T2
1
T1 1
P2
P1
(12.15)
361
362
Fuel Cell Systems Explained
By rearranging equation (12.15), the change in the temperature is given by:
1
T
T2 T1
c
T1
P2
P1
1
(12.16)
Note that because P2 < P1, ΔT will always be negative. This expression enables the
derivation of a formula for the power available from the turbine. Applying the same
reasoning and simplifying the assumptions as undertaken in Section 12.4.2 yields:
1
Power C P
C P cT1
T m
P2
P1
1
m
(12.17)
To obtain the power available to drive an external load, the power should be
multiplied by the mechanical efficiency which, as for compressors, should be 0.98 or
above. Turbine performance can be represented using charts in exactly the same way
as for compressors, except that in the case of turbines the vertical axis becomes
P1/P2 instead of P2/P1 and the direction of the RSF lines is completely different. An
example of a chart for a radial turbine is presented in Figure 12.8. For any given
turbine speed, the mass flow rate rises as the pressure drop increases, as would be
expected, but tends towards a maximum value, which is called the ‘choking limit’.
Naturally enough, the value of the choking limit depends largely on the diameter of
the turbine housing.
Worked examples of compressor and turbine calculations are given in Appendix 3.
12.1.3
Ejector Circulators
With no moving parts, the ejector is the simplest of all types of pump. In a fuel cell that
operates with gaseous fuels stored at pressure, an ejector harnesses the incorporated
mechanical energy to circulate the given fuel through the stack. The device is widely
chosen to perform this function in hydrogen fuel‐cell systems and some SOFCs.
A diagram of a simple ejector is shown in Figure 12.9. A gas or liquid passes through
the narrow pipe A and enters the venturi B. It acquires a high velocity at B and hence
produces suction in pipe C. The fluid passing through A thus entrains the fluid from C
and discharges it at D. The fluid from A, which must be at a higher pressure than that in
C, B or D, does not have to be the same as that in pipe C, B or D. Ejectors are frequently
found in steam systems, with steam being the fluid passing through the narrow pipe and
jet A. Ejectors can also be used to pump air, to maintain a vacuum in the condensers of
steam turbines or to pump water into boilers.
In a hydrogen fuel‐cell system, an ejector circulates the fuel gas through the stack.
Hydrogen is supplied at high pressure at A, and the energy of the expanding gas in the
ejector draws in gas from the anode exhaust at C and, together with the fresh hydrogen,
sends it on through D to the anode inlet. The pressure difference generated will be
sufficient to drive the gas through the cell as well as through any humidification
equipment in the case of a PEMFC. The internal diameters of pipes A, C and D and the
The Complete System and Its Future
η = 0.70
4.0
3.5
2.5
6000
η=0
.65
Rotor speed
factor/
rev. min–1 K–½
η = 0.
55
η=0
.60
Pressure ratio
3.0
5000
2.0
4000
1.5
3000
1.0
0
0.25
0.5
0.75
Mass flow factor/kg
s–1
K½
1.0
1.25
bar–1
Figure 12.8 Performance chart for a typical small radial turbine.
A
B
D
C
Figure 12.9 Diagram of a simple ejector circulation pump.
mixing region B to suit the pressure differences and flow rates associated with the
required duty can be obtained from chemical engineering reference books.
12.1.4
Fans and Blowers
Fans or blowers deliver straightforward air cooling in a wide range of equipment from
desktop computers to cars. The axial fan, which is employed for cooling electronic
363
364
Fuel Cell Systems Explained
equipment, is an excellent device for moving air but provides only a very small pressure
lift. For example, air blown at a rate of about 0.1 kg s−1 by a small axial fan may drop to
zero if the back‐pressure rises even to 50 Pa (0.5 cm water pressure). Nevertheless, axial
fans have been included in a few open designs of PEMFC.
By contrast, the centrifugal fan, which blows air through air‐conditioning systems,
provides a somewhat greater pressure lift. This type of fan is not too dissimilar to the
centrifugal compressors described earlier, except that it runs at much lower speeds (by
a factor of several hundred), has much longer blades and a considerably more open
construction. There are various types of centrifugal fan that differ in the design and
orientation of the blades.
Blowers and fans are normally employed to assist in the removal of heat (cooling)
from the gas stream rather than increasing the kinetic energy of the gas. In a cooling
system, the effectiveness is a function not only of the power consumed by the fan
or blower but also of the design of heat exchanger employed to cool the gas.
Rather than using the term efficiency in such cases, it is more practical to define
effectiveness as:
rate of heat removal kWh h
Cooling system effectiveness
1
(12.18)
electrical power consumed (kW)
As an example, consider a small axial fan with a 120‐mm diameter that is often used
to cool electrical equipment. Such a fan might move air at a rate of 0.084 kg s−1 and
consume 15 W of electrical power. If the air it blows rises in temperature by 10°C, then
the rate of removal of heat will be given by:
1004 10 0.084 843 W
Power C P DT m
(12.19)
That is, 843 W of heat is removed for just 15 W of electrical power, so the cooling system
effectiveness is 843/15 = 56.
There is always a balance, or trade‐off, in cooling systems between the flow rate of air
and the electrical power consumed. Higher flow rates improve heat transfer, but at the
expense of more power consumed by the fan.
12.1.5
Pumps
The blowers and fans considered earlier are best suited to high flow rates of gas
with very small pressure lift. With small‐to‐medium‐sized PEMFC stacks of, say,
200 W to 2 kW, the back‐pressure on both the air and fuel is too high for blowers
and fans. On the other hand, the pressures and flow rates will be too low for any of
the commercial compressors that are discussed earlier. For small PEMFC systems
another type of pump is required, the main features of which should include the
following:
●
●
●
Low cost.
Silent.
Reliable long‐term operation.
The Complete System and Its Future
Figure 12.10 Diagram of a diaphragm
pump — a device with the advantages
of low‐cost, quiet operation and
long‐term reliability.
Soft rubber
diaphragm
Soft rubber
valves
●
●
Available in a range of sizes that will also accommodate small‐scale PEMFC systems,
e.g., gas flow rates from about 2.5 × 10−4 to 2.5 × 10−3 kg s−1, i.e., 12–120 standard litre
per minute (SLM).2
Efficient, low‐power consumption.
The most suitable pumps that address these requirements are either small vane or
diaphragm designs. The diaphragm pump, as commonly employed in fish tanks, meets
most of the earlier requirements. There are many variations in design, but the basic
operating principle can be understood from Figure 12.10. The diaphragm is moved up
and down by an electrical motor to shift the air through the system by way of two valves
in an obvious mechanical manner. The diaphragm is made from soft rubber and thereby
provides quiet operation over long periods, although the pressure lift is limited to
10–20 kPa. The flow rate can be controlled easily by modulating the force applied by the
pump actuator. For instance, it has been reported that a 300‐W PEMFC has employed
a diaphragm pump to supply a required reactant air flow of 10–20 SLM. This flow rate
is delivered at a pressure of between 110 and 115 kPa, and the motor–pump combination
consumes between 14 and 19 W. The parasitic system power loss due to the air pump is
thus about 6%, which is acceptable for such a small system.
12.2
Power Electronics
The electrical output of a stack usually needs to be conditioned to match the demands
of the particular application. Some operations require a constant or near‐constant
voltage and others a conversion of the DC output to AC. Power electronic components
2 The standard litre per minute (SLM) is a unit of volumetric flow rate of a gas corrected to ‘standardized’
conditions of temperature and pressure, which can vary considerably between different fields of science and
engineering.
365
Fuel Cell Systems Explained
can be added to achieve these requirements. A voltage regulator is used to stabilize the
DC voltage, and an inverter is employed to convert DC into AC. These items of power
electronics are described in the following subsections.
12.2.1
DC Regulators (Converters) and Electronic Switches
As shown in Figures 3.1 and 3.2 in Chapter 3, the voltage from a fuel cell falls with
increasing current density. By way of example, data from a 250‐kW PEM fuel‐cell system
that powered a bus3 are presented in Figure 12.11. The stack voltage varied from about
400 to over 750 V during operation and had different values at the same current.
The variability arises because, as well as current, the voltage is also dependent on
many other factors such as operating temperature and air pressure. Such behaviour is
not compatible with most electronic and electrical devices as they normally require a
power supply with a fairly constant voltage. In most fuel‐cell systems, it is therefore
necessary to stabilize the voltage supplied from the stack to the electrical equipment,
either by dropping the voltage down to a fixed value below the operating range of
the fuel cell or by boosting it up to a fixed value above the operating range. Changes
in DC voltage are achieved by using ‘switching’ or ‘chopping’ circuits, which are
described in the following text. These circuits, as well as the inverters discussed in
Subsection 12.2.4.2, operate with electronic switches. Although the particular type of
electronic switch is not of great concern, it is useful to outline the main characteristics
800
Depending on conditions, the
voltage/current points can be
anywhere in this region
600
Stack voltage (V)
366
400
200
50
100
150
200
Stack current (A)
Figure 12.11 Data from a 250‐kW fuel‐cell system designed to power a bus. (Source: Derived from data
in Spiegel, RJ, Gilchrist, T and House, DE, 1999, Fuel cell bus operation at high altitude, Proceedings of
the Institution of Mechanical Engineers, Part A, vol. 213, pp. 57–68. Reproduced with permission of Sage
Publications.)
3 From: Spiegel, RJ, Gilchrist, T and House, DE, 1999, Fuel cell bus operation at high altitude, Proceedings of
the Institution of Mechanical Engineers, Part A, vol. 213, pp. 57–68.
The Complete System and Its Future
Table 12.1 Key data for the main types of electronic switch used in power electronics.
Type
Thyristor
MOSFET
Symbol
IGBT
d
c
g
g
s
e
Maximum voltage (V)
4500
1000
1700
Maximum current (A)
4000
50
600
Switching time (µs)
10–0.5
0.3–0.5
1–4
IGBT, insulated‐gate bipolar transistor; MOSFET, metal-oxide-semiconductor field‐effect
transistor.
of those that are most commonly employed, as given in Table 12.1, and to discuss their
respective advantages and disadvantages.
The metal-oxide-semiconductor field‐effect transistor (MOSFET) is a solid‐state
switch that is turned on by applying a voltage, usually between 5 and 10 V, to the gate ‘g’
shown in symbol displayed in Table 12.1. When in the ‘on’ or ‘closed’ state, the resistance
between the drain ‘d’ and the source ‘s’ is very low. The power required to ensure a very low
resistance is small, as the gate current is low. The gate does, however, have a considerable
capacitance, so special drive circuits are usually required. The current path behaves like
a resistor, with a certain built‐in value of the transistor’s internal resistance, RDS,on, in the
‘on’ state. In voltage regulation circuits, the value of RDS,on for a MOSFET can be as low as
0.01 Ω. Such low values are only possible with devices that can switch low voltages, in the
region of up to 50 V. Devices that can switch higher voltages have RDS,on values of about
0.1 Ω that lead to higher power losses. Consequently, MOSFETS are generally employed
in low‐voltage power electronics systems of less than about 1 kW.
The integrated gate bipolar transistor (IGBT) is essentially a three‐terminal integrated
circuit that combines a conventional bipolar transistor and a MOSFET — therefore, it has
the advantages of both. A low voltage with negligible current is applied to the gate to
switch it to the ‘on’ state. The main current flow is from the collector ‘c’ to the emitter ‘e’, as
indicated by the symbol in Table 12.1, and this path has the characteristics of a p–n
junction. The result is that the voltage does not increase much above 0.6 V at all currents
within the rating of the device. This feature makes the IGBT the preferred choice for systems
in which the current is greater than about 50 A. The IGBT can also be made to withstand
higher voltages. The longer switching times compared with the MOSFET, as given in
Table 12.1, are a disadvantage in low‐power systems. Nevertheless, the IGBT is now almost
universally the electronic switch of choice in systems from 1 kW up to several hundred kW,
with the ‘upper’ limit rising each year with ongoing improvement in the technology.
The thyristor has been the electronic switch most commonly adopted in small‐scale
power electronics such as dimmer switches for electric lights. Unlike the MOSFET and the
IGBT, the thyristor can only serve as an electronic switch — it has no other application.
The transition from the ‘off’ to the ‘on’ state is triggered by a pulse of current into the gate.
The device then remains in the ‘on’ or conducting state until the current flowing through
it falls to zero. This characteristic makes the thyristor particularly useful in circuits
367
368
Fuel Cell Systems Explained
Figure 12.12 Symbol used for an electronically operated switch.
dedicated to the rectification of AC. Various types of thyristor, in particular the gate turn‐
off (GTO) version, can be switched off even when current is flowing by the application of
a negative current pulse to the gate. Despite the fact that the switching is achieved by just
a pulse of current, the energy required to effect the switching is much greater for the
thyristor than for the MOSFET or the IGBT. Furthermore, the switching times are
markedly longer. The only advantage shown by the thyristor (in its various forms) for
DC switching is that higher currents and voltages can be accommodated.
The circuit symbol for a switching device, whether MOSFET, IGBT or thyristor, is
generally the ‘device independent’ symbol shown in Figure 12.12. In all cases, it is essential
that the switch moves as quickly as possible from the conducting to the blocking state, and
vice versa. No energy is dissipated in the switch when it is open and only very little energy
is lost when it is fully closed; it is during the transition between open and closed states that
the product of voltage and current is non‐zero and power is lost. This latter loss appears as
heat and therefore in all cases some cooling of the switching device is required, and is
usually provided by mounting it on a heat sink of large area that may be cooled with an air
blower.
12.2.2
Step‐Down Regulators
The electronic switches that have just been described are used in regulators such as the
switch‐mode ‘step‐down’ or ‘buck’ switching regulator (or chopper), as shown diagrammatically in Figure 12.13. The essential components are an electronic switch with an
associated drive circuit, a diode and an inductor. When the switch is on, the current
from the fuel‐cell stack flows through the inductor and the load, as indicated by the
circuit in Figure 12.13a. A voltage drop across the inductor, due to the induced magnetic
field inside the coil, initially limits the current through the circuit.4 The current gradually
rises over a short time as the inductor becomes fully magnetized. The switch is then
turned off, and the energy stored in the inductor keeps the current flowing through the
load via the diode, as illustrated by the circuit in Figure 12.13b. The different currents
during each part of this on–off cycle are shown in Figure 12.14. If necessary, the voltage
across the load can be further smoothed through the use of capacitors.
4 The effect is often described as a ‘back‐voltage’, because it is the voltage that ‘pushes back’ against the
current that induces it. The back‐voltage is the voltage drop in an AC circuit caused by magnetic induction.
The Complete System and Its Future
Inductor
(a)
V1
Fuel cell
On
Load
Current path when switch is on
Inductor
(b)
V1
Fuel cell
Load
Off
Current path when switch is off
Figure 12.13 Circuit diagrams showing the operation of a switch‐mode, step‐down, voltage regulator:
(a) current path when switch is on and (b) current path when switch is off.
On
On
On
Off
Off
Off
Current supplied by fuel cell during the on time
Current circulating through diode during the off time
Current through load, being the sum of these two components
Figure 12.14 Currents in the switch‐mode, step‐down, regulator circuit.
369
370
Fuel Cell Systems Explained
If V1 is the supply voltage and the ‘on’ and ‘off ’ times for the electronic switch are ton
and toff, then it can be shown that the output voltage, i.e., the voltage across the load, V2,
is given by:
V2
ton
ton toff
V1
(12.20)
The regular variations in voltage across the load are known as ripple, and these are
influenced by the frequency of switching — at higher frequency, the ripple is less.
Nonetheless, each turn‐on and turn‐off involves the loss of some energy, and therefore
the frequency should not be too high. A control circuit is needed to adjust ton to
achieve the desired output voltage, and such circuits are readily available from
manufacturers.
The main energy losses in the step‐down chopper circuit are as follows:
●
●
●
●
Switching losses in the electronic switch.
Power lost in the switch while it is turned on (0.6 × I for an IGBT or RDS,on × I2 for a
MOSFET).
Power lost because of the resistance of the inductor.
Losses in the diode, 0.6 × I where I is the current in the circuit and the value of 0.6 is
taken to be the voltage loss over the diode.
In practice, all of these losses can be reduced to very low levels. The efficiency of such a
step‐down chopper circuit should be at least 90% and in systems with voltages of 100 V
or higher, efficiencies of over 98% are routinely achieved.
As a note of caution, mention should be made of an alternative to the step‐down
regulator, namely, the ‘linear’ regulator circuit. A transistor that has been designed to
provide a variable resistance between the emitter and the collector, rather than an on–
off function, is employed in this circuit. The gate voltage is adjusted so that the device
resistance is at the correct value required to drop the cell voltage. The linear circuit is
commonly used in some low‐cost electronic devices but is inefficient and is a poor
choice for fuel‐cell systems.
12.2.3
Step‐Up Regulators
Since fuel cells are low‐voltage devices, it is usually desirable to step up or boost the DC
voltage. This can also be simply and efficiently done by using electronic‐switching
circuits. The circuit of a typical switch‐mode, step‐up voltage regulator is shown in
Figure 12.15, and its operation is as follows. The starting assumption is that there is
charge in the capacitor. When the switch is on, an electric current builds up in the
inductor, as shown in Figure 12.15a. The load is supplied by the discharge of the capacitor.
The diode prevents the charge from the capacitor flowing back through the switch.
When the switch is off, as in Figure 12.15b, the voltage across the inductor increases
sharply because the current is falling. Once the voltage rises above that of the capacitor
(plus about 0.6 V for the diode), the current will flow through the diode, charge up the
capacitor and pass through the load. This will continue as long as there is still energy in
the inductor. The switch is then closed again.
The Complete System and Its Future
(a)
Inductor
V1
Fuel cell
On
Load
Current paths while switch is on
(b)
Inductor
V1
Fuel cell
Off
Load
Current flow while switch is off
Figure 12.15 Circuit diagram to show the operation of a switch‐mode, step‐up, voltage regulator:
(a) current path when switch is on and (b) current path when switch is off.
High voltages across the load are achieved by having the switch turned off for a
shorter period than when in the ‘on’ position. For an ideal switch‐mode regulator with
no losses, the voltage across the load (V2) is given by:
V2
ton toff
toff
V1
(12.21)
where toff and toff, as before, are the times when the switch is on or off.
In practice, the output voltage will be somewhat less, for similar reasons as those for
the step‐down regulator. Control circuits are readily available from many manufacturers
for both step‐down and step‐up voltage regulators.
The losses in the step‐up regulator arise from the same sources as in the step‐down
regulator. Since, however, the currents through the inductor and switch are higher than the
current through the load, the energy losses are also higher. Also, as all the charge passes
through the diode, there will be an energy loss associated with this current. Given these
disadvantages, step‐up converters are generally less efficient than step‐down converters,
although values of 95% or more are achievable. The step‐up and step‐down switching or
chopping circuits are generally referred to as ‘DC–DC converters’, and commercial packages
are readily available covering a wide range of power and voltage requirements.
12.2.4
Inverters
Fuel‐cell systems that are designed to supply electricity to homes and businesses need
to generate AC power. For small, single dwellings, single‐phase AC is required, whereas
for more substantial installations a three‐phase supply is generally required. In some
371
372
Fuel Cell Systems Explained
cases, the fuel‐cell system may be operating in parallel with a conventional power
supply grid; in other situations, it may be a stand‐alone system or grid independent. The
rapid emergence of solar photovoltaic (PV) systems in recent years has encouraged
power electronics companies to develop inverters that will accept DC voltages from
solar panels, and many different types and classes of inverter are on the market.
Therefore, given the considerable body of literature that is now available, only a basic
description of the operating principle of an inverter will be presented here.
12.2.4.1
Single Phase
The basic circuit diagram for a single‐phase inverter is shown in Figure 12.16. There
are four electronic switches, labelled A, B, C and D, that are connected in a so‐called
H‐bridge. Across each switch is a diode. A resistor and an inductor represent the load
through which the AC is to be driven. The circuit operates as follows. First, switches
A and D are turned on (B and C are open) and a current flows to the right through the
load. These two switches are then turned off, and switches B and C are turned on, thereby
causing a current to flow in the opposite direction through the load, i.e., right to left. The
diodes across the four switches provide a safe path for any charge to dissipate when
the switches are turned off. The resulting current waveform is shown in Figure 12.17. In
some situations, though increasingly few, this waveform will be adequate.
Most homes and businesses are currently supplied with electricity, which is generated in
thermal power stations by means of rotating equipment that produces AC of varying voltage in the form of an almost perfect sine wave at a fundamental frequency of 50 or 60 Hz.5
A
B
Fuel cell
Load
C
D
Figure 12.16 H‐bridge inverter circuit for producing single‐phase alternating current.
5 The frequency of distributed power was determined by the rotation of steam‐powered generators at the
end of the 19th century. In the early days of electrification, so many frequencies were used that not one value
prevailed (London in 1918 had 10 different frequencies). The proliferation of frequencies grew out of the
rapid development of electrical machines in the period 1880–1900. In the early incandescent lighting period,
single‐phase AC was common and typical generators were 8‐pole machines operated at 2000 rpm that gave a
frequency of 133 cycles per second. As distribution networks developed, it became necessary to standardize
the frequency. In Europe, generator manufacturers chose 50 Hz, whereas, for the most part, manufacturers in
the United States adopted the 60 Hz standard.
The Complete System and Its Future
Current
in load
A and D on
B and C on
A and D on
Time
Figure 12.17 Current versus time graph for a square‐wave, switched mode, single‐phase inverter.
By contrast, the more square waveform delivered by the simple circuit of Figure 12.16 is
incompatible with distributed electricity and the appliances that it powers. The
incompatibility arises because a square voltage waveform is composed of many different
frequencies, or harmonics, in addition to the fundamental frequency. The higher
frequency harmonics can have harmful effects on other equipment connected to the
distribution network (grid), and on cables, transformers and switchgear. Harmonics
can cause inefficiencies in electric motors and damage to computers and other electronic
equipment. The possibility of such adverse behaviour now requires networks to impose
strict regulations concerning the ‘purity’ of the waveform of any AC power source that
may be connected to the grid. The standards vary between countries and inverter
manufacturers have to ensure that the AC output is as near to a pure sine wave as
possible with a minimum of harmonics, or ‘harmonic distortion’.
To achieve purity in the AC generated by an inverter, pulse width modulation
(PWM) is employed to control the IGBTs or other electronic switches employed.
More recently the ‘tolerance band’ technique has been applied (v.s.). The principle of
PWM is shown in Figure 12.18 and relates to the same circuit as shown in Figure 12.16.
In the positive cycle, only switch D is on all the time, while switch A is active only
intermittently (i.e., pulsed on–off ). When A is on, current starts to flow through the
load and switch D. The load inductance causes a back‐voltage, which ensures that the
current is initially small and increases over a very short time. When A is turned off,
the current that has built up continues to flow around the bottom right loop of the
circuit due to the load inductance through switch D and the ‘free‐wheeling’ diode in
parallel with switch C.
A similar process occurs in the negative cycle, except that switch B is on all the time
and switch C is ‘pulsed’. When C is on, current builds up in the load, and when C is off,
it continues to flow (though declining) through the upper loop in the circuit and through
the diode in parallel with switch A. Control of switches A and C is carried out by an
electronic circuit that modulates the on–off times, i.e., pulse widths according to a pre‐
determined sequence to generate a varying voltage that best approaches that of a sine
wave. The precise shape of the voltage waveform generated over the load will depend on
the nature (resistance, inductance and capacitance) of the load as well as the pulsing of
switches A and C, but a typical half‐cycle is shown in Figure 12.19. The waveform is still
373
374
Fuel Cell Systems Explained
Time
Sinusoidal current required
Control voltages on electronic switches of H bridge
A
D
C
B
Figure 12.18 Pulse‐width modulation switching sequence for producing an approximately sinusoidal
alternating current from the circuit of Figure 12.17.
Voltage
Time
Figure 12.19 Typical voltage versus time waveforms from a pulse‐modulated inverter.
The Complete System and Its Future
not a sine wave but is much closer than that of Figure 12.17. Clearly, the more pulses
there are in each cycle, the closer will be the wave to a pure sine wave and the weaker
will be the harmonics. A common standard is 12 pulses per cycle. This requires the
switches to be able to operate at frequencies twelve times those of normal a 50 Hz
supply, i.e., at least 600 Hz.
One of the issues with PWM is that to improve the sine wave output, i.e., limit the
harmonics, high frequency switching is required. This can place a strain on
the electronic switches and can lead to low inverter efficiency. In modern inverters,
the switching pulses are generated by microprocessor circuits, and some very
sophisticated approaches have been developed that are outside the scope of this
book. One example is the abovementioned tolerance band pulse method, in which
switching is controlled so that the output voltage at any time mirrors that of a waveform
generated by the microprocessor circuit, within a certain tolerance band of upper
and lower voltage limits. The method is illustrated in Figure 12.20. The output voltage
is continuously monitored and compared with an internal ‘upper limit’ and ‘lower
limit’, which are sinusoidal functions of time. In the positive cycle switch D
(in Figure 12.16) is on all the time. Switch A is turned on, and the current through the
load rises. When it reaches the upper limit, A is turned off, and the current flows on,
though declining, through the diode in parallel with C, as before. When the lower
limit is reached, switch A is turned on again, and the current begins to build up again.
This process is continuously repeated, with the voltage rising and falling within the
tolerance band.
In Figure 12.20, the on/off cycle is shown in (a) for a wide tolerance band and in (b) for
a narrow tolerance band. It should be appreciated that the resistance and inductance of
the load will also affect the waveform and hence the frequency at which switching
occurs. The method is therefore an adaptive system that always keeps the same deviation
from a sine wave and hence limits the unwanted harmonics below fixed levels. Tolerance
band methods are employed in multilevel inverters that are built to raise AC voltages to
high levels for power distribution on networks.
(a)
(b)
C
On
C
Off
C
On
On
On
Off
Upper limit
On
Off
Off
Upper limit
Lower limit
Lower limit
Figure 12.20 Graph summarizing some data from a real 250‐kW fuel‐cell system used to power a bus:
(a) narrow tolerance band and (b) wide tolerance band. (Source: Derived from data in Spiegel, RJ,
Gilchrist, T and House, DE, 1999, Fuel cell bus operation at high altitude, Proceedings of the Institution
of Mechanical Engineers, Part A, 213, 57–58.)
375
376
Fuel Cell Systems Explained
12.2.4.2 Three Phase
In almost all parts of the world, AC electricity is generated and distributed using three
parallel circuits, the voltage in each one being out of phase with the next by 120°. While
most homes are supplied with just one phase, most industrial establishments have all
three phases available. For industrial combined heat and power (CHP) systems, for
example, the DC from the fuel cell will need to be converted to three‐phase AC.
A three‐phase inverter is only a little more complicated than the single‐phase device.
The basic circuit is illustrated in Figure 12.21a. The output of the inverter is shown
connected to the primary of a three‐phase transformer. Six switches, with free‐wheeling
diodes, are connected to the primary windings of a three‐phase transformer on the
right, which represents the load. Equally, the load could be a three‐phase motor such
as would be found on an industrial pump or compressor, each of which could be BoP
components in a MW‐scale fuel‐cell system. Adopting the same principle described for
the single‐phase circuit of Figure 12.17, the switches enable the generation of three
(a)
Three-phase
transformer
primary
(b)
Voltage
A
B
C
Time
Time
Time
Figure 12.21 (a) Circuit diagram of a simple three‐phase DC–AC inverter and (b) current versus
time graph for the inverter assuming a purely resistive load. One complete cycle for each phase is
shown. Current flowing out from the common point is taken as positive.
The Complete System and Its Future
similar, but out‐of‐phase, voltage waveforms. Each cycle can be divided into six steps.
The graphs in Figure 12.21b show how the current in each of the three phases changes
with time through employing this simple arrangement. As with the single‐phase inverters
described earlier, the switching sequence in a three‐phase inverter is modified — by
means of pulse‐width modulation or tolerance band methods — to achieve current and
voltage waveforms that closely approach that of a sine wave at the required frequency.
The modern three‐phase ‘Universal’ inverter is built along similar lines whether it is
for high or low power, and whether it is ‘line‐commutated’ (i.e., the timing signals are
derived from the grid to which it is connected) or ‘self‐commutated’ (i.e., independent
of the grid). Indeed, the same basic circuit is used irrespective of the modulation
method. Particularly with the growth in solar PV, both single‐phase and three‐phase
inverters have become commodity items. The circuit is shown in Figure 12.21(a). Signals
to turn the switches on and off are taken from a microprocessor. Voltage‐ and current‐
sensing signals may be taken from the three phases, the input, each switch or other
places. Digital signals from sensors may also be employed and both instructions and
information may be sent to and received from various parts of the system. In all cases
the hardware will essentially be the same, i.e., as shown in Figure 12.22. Inverter units
have thus become like many other electronic systems — a standard piece of hardware
that can be programmed for diverse applications.
Analogue-toVoltage
digital
and current
converter
sense signals
Microprocessor
Digital
signals
e.g., alarms
Data to and
from a
supervisory
controller
Figure 12.22 ‘Universal’ three‐phase inverter.
To gates
of the six
electronic
switches
377
378
Fuel Cell Systems Explained
12.2.5
Fuel‐Cell Interface and Grid Connection Issues
The point has been made throughout this book that the power output of a fuel‐cell
stack, as measured by the voltage and current, is dependent on various operating
parameters such as fuel and oxidant flow rates, pressure and temperature. The output
of the stack, in turn, impacts on the DC–DC converter, which may also influence the
inverter. Unlike the inverter in a power generation system such as a PV array or wind
turbine, however, the performance of the inverter in a fuel‐cell system can actually
affect the fuel‐cell stack. The main issue is that excessive ripple currents, which may be
caused by faulty PWM in the inverter, are known to cause degradation of the fuel‐cell
catalysts. Provision must therefore be made for the monitoring of such currents and the
annunciation of an alarm if the levels get too high.
When designing a fuel‐cell system for stationary application, e.g., cogeneration
systems from a few kW up to many MW, the power that is generated has to meet certain
quality standards before the facility can be connected to the electricity distribution grid.
Again, an important issue is the level of harmonics generated in the power electronics
circuits. Devices must also be in place to protect the fuel cell, the inverter and the grid
from faults such as short-circuits (surge protection) and lightning strikes. Such faults
can develop with any power generation equipment (e.g., privately owned solar PV or
wind turbines) and are not specific to fuel‐cell systems. Any grid‐connected inverter
also has to produce AC power that matches the existing power presented on the grid. In
particular, a grid‐interactive inverter must match the voltage, frequency and phase of
the power line to which it is connected. There are numerous technical requirements to
the accuracy of this tracking.
Home and business owners that have installed solar PV arrays may be able to sell
excess power from their system back into the grid, i.e., when the production exceeds the
demands of the home or business load. For PV systems, this situation may occur in the
middle of the day when solar incidence is at its highest. Utilities have set feed‐in tariffs
that encourage such flow back into the grid. Given the right circumstances, power
returned to the grid from a distributed generator (DG) such as a fuel‐cell system can
reduce the need for upgrading the local grid. Indeed, although a fuel cell may be sized
to meet the load of the home or business, there may be occasions when it becomes
profitable for excess power generated locally to be exported into the grid. Grid‐
connected systems also benefit in that if the DG fails, the grid provides emergency
backup. If, however, the grid fails, e.g., through a lightning strike at some location on the
high‐voltage transmission system, a situation can arise in which a DG continues to
power a section of the distribution system. The incident is known as ‘islanding’ and can
be dangerous to utility workers, who may not realize that a circuit is still powered.
When an islanded fuel cell reconnects to the grid, it is essential that the inverter has
provision to lock the phase of its AC voltage waveform to that of the grid.
12.2.6
Power Factor and Power Factor Correction
The voltage and current of an AC circuit are sinusoidal in nature (see Figure 12.23), i.e.,
the amplitude of both the current and the voltage of an AC circuit constantly changes
over time. In a purely resistive AC circuit, voltage and current waveforms are in step
(or in phase) and thus change polarity at the same instant in each cycle. All the power
The Complete System and Its Future
Voltage
Current
Figure 12.23 Voltage and current out of phase. The reactive power can be generated locally by
distribution systems such as fuel cells.
VA
r/k
nt
re
pa
e
ow
p
Ap
Reactive power/kVAR
Phase angle (ϕ)
Real or true power/kW
Figure 12.24 Relationship between true power, apparent power and reactive power.
entering the load is consumed (or dissipated). Such power is defined as the ‘true power’,
also known as ‘active power’, ‘real power’ or ‘useful power’. This ideal situation is rarely
met in practice because loads are reactive, i.e., they have elements such as capacitors or
inductors present, and consequently create a phase difference between the current and
voltage waveforms. During each cycle of the AC voltage, extra energy — in addition to
any energy consumed in the load — is temporarily stored in the load and then returned
to the grid in a fraction of the cycle later. The extra energy is known as the ‘reactive
power’. The combination of reactive power and true power gives the ‘apparent power’,
which, in an AC circuit, is the total of all the power both dissipated in the load and
absorbed or returned to the grid. When the circuit is purely resistive, the apparent power
is equal to true power, but in an inductive or capacitive circuit, the apparent power is
greater than the true power. Therefore, the following are encountered in AC circuits:
●
●
●
True (active, real or useful) power, expressed in watts (kW).
Reactive power, usually expressed in reactive volt‐amperes (kvar).
Apparent power, usually expressed in volt‐amperes (kVA).
To help understand the relationship between true power and reactive power, the
two parameters are shown as two sides of a right‐angled triangle, as depicted in
Figure 12.24. The angle Ф describes the phase shift between the voltage and current.
The larger the phase angle, the greater the reactive power is generated by the system.
The hypotenuse of the triangle is the apparent power, i.e., the vector sum of the
379
380
Fuel Cell Systems Explained
true power and the reactive power. The ratio between the true power and the apparent
power is known as the ‘power factor’ (PF) and is the cosine of Ф.
The power factor is a dimensionless number between −1 and 1. When PF = 0, the
energy flow is entirely reactive and stored energy in the load returns to the source on
each cycle. When the power factor is 1, all the energy supplied by the source is
consumed by the load (i.e., the load behaves as a pure resistor with no capacitance or
inductance). A power factor of −1 can result from returning power to the grid, such as
in the case of a building fitted with solar panels when their power is not being fully
utilized and the surplus is fed back into the supply network. Power factors are usually
stated as ‘leading’ or ‘lagging’ to show the sign of the phase angle. Capacitive loads are
leading (current leads voltage), and inductive loads are lagging (current lags voltage).
A low‐power factor (for which the consumer may be penalized by the network operator)
can be increased to an acceptable level (usually above 0.9) by installing a power factor
correction (PFC) unit close to the load. The most widely adopted of the PFC units are
passive and comprise a series of capacitors, which are incrementally brought on stream
by contactors that engage as the load changes. Power factor correction can also be
achieved with an active power electronics system that incorporates inverter technology,
as described earlier. The switching of the pulses in inverter circuits can be programmed
by the user, and in most modern inverters the phase angle can be adjusted when the
inverter is installed. This feature is particularly beneficial for fuel‐cell systems since they
can often be installed close to the load, unlike other distributed power generation
systems such as large‐scale solar or wind farms.
12.3
Hybrid Fuel‐Cell + Battery Systems
In general, all types of fuel‐cell stack perform best if they are run under a constant load.
Large fluctuating cell voltages can accelerate degradation of both the anode and cathode
catalysts and thereby reduce stack lifetime. Combining a rechargeable battery or supercapacitor6 with a fuel cell creates a hybrid system that can accommodate the variations
in load that occur during service. When the total power requirements from a hybrid
system are low, the surplus electrical energy is stored in the rechargeable battery or
capacitor. Conversely, when the power demand exceeds that available from the fuel cell,
energy is taken from the battery or capacitor. In essence, hybrid systems are using the
fuel cell as a battery charger.
Two extreme types of hybrid may be considered. One of these facilities relates to a
situation where the electrical power demand varies in a predictable manner and,
therefore, the hybrid system is mainly in ‘standby’ mode for fairly long periods, and the
fuel cell serves to recharge the battery. Whereas the battery is required to supply most
of the power during ‘transmit’ or peak periods, the fuel cell operates more or less
continuously in providing average power to charge the battery. In turn, the battery
must deliver sufficient power, hold enough energy for the regular and frequent loads
6 The advantages of the capacitor are that the charge–discharge cycle is more efficient and much faster. The
disadvantages are that capacitors store much less energy for a given space and are more expensive per watt‐
hour stored.
The Complete System and Its Future
DC–DC inverter
Rechargeable
battery
Fuel-cell
stack
L
o
a
d
EN
Battery fully
charged
sense
Figure 12.25 Diagram of a simple fuel‐cell–battery hybrid system.
and be conducive to recharging in the periods in between. Examples of this set of
requirements are to be found with certain data‐logging devices, telecommunications
systems and land‐ or buoy‐based navigation equipment. The other extreme type of
hybrid is designed for situations in which the power demand can be highly irregular
and unpredictable as encountered, for instance, with mobile phones. In most cases,
the combination of battery and fuel cell offers the additional benefit of a lower cost
compared with a fuel‐cell system that is sized to meet the highest demand.
A hybrid concept that employs a fuel cell and a battery is shown schematically in
Figure 12.25. In addition to the components shown, a controller is required to prevent
overcharging of the battery. As discussed in Section 12.2.3, a DC–DC converter may be
required to match the output voltage of the fuel‐cell stack with that of the battery/load.
Such a system would be attractive for application with direct methanol fuel cells
(DMFCs) or other cells that run on liquid fuels, from all of which the average power is
very low. Full‐hybrid arrangements (also known as ‘hard hybrids’) are those in which the
peak power that may be required is much greater than the average power that can
be delivered by the fuel cell and therefore takes full advantage of power that is stored in
the battery.
By contrast, a mild (or ‘soft’)‐hybrid system is one in which the battery power and
energy storage are quite low compared with the power delivered by the fuel cell. Such
systems may be found in vehicles and boats, for example, where there is little difference
between the peak and average power demands. An example of the variation in system
power with time for a mild hybrid is shown in Figure 12.26a. The fuel‐cell power is
sufficient for most of the time, but the battery can ‘shave’ or ‘lop’ the peaks off the power
requirement and thus can substantially reduce the required fuel‐cell capacity. During
times when the fuel‐cell power exceeds the power demand, the battery is recharged.
This profile is typical of urban electric vehicles, where the peaks correspond to occasions
such as accelerating from traffic lights, but for most of the time the vehicle will be
proceeding slowly and steadily, or else it will be stationary.
An additional option is illustrated in the power–time graph of Figure 12.26b. Here
an electric vehicle is using the motor as a brake. The electric motor functions as a
generator to convert the motion energy into electrical energy. The energy can either
be passed through a resistor and dissipated as heat (referred to as ‘dynamic braking’)
or can be passed to a rechargeable battery, to be used later to run the motor. The latter
381
Fuel Cell Systems Explained
Power
(a)
Average power
Time
(b)
Maximum
fuel-cell power
Power
382
Braking
Time
Figure 12.26 Power versus time graphs for systems suitable for a soft hybrid electric vehicle:
(a) without and (b) with regenerative braking.
method is known as ‘regenerative braking’ and is clearly the better option from the
point of view of system efficiency, but it does presuppose a hybrid system with a
rechargeable battery.
A fairly sophisticated control system with a responsive fuel cell is required when regenerative braking is incorporated in a vehicle. The battery needs to be operated in a partial
state‐of‐charge condition so that it can absorb the electrical energy supplied by the motor
during braking. The flow of power from the fuel cell to the battery will have to be properly
controlled, since the duty cycle of the vehicle is much more variable than one without
regenerative braking. That is, the fuel cell needs to respond quickly depending on whether
it is required to provide a small amount of power to top up the battery or whether a burst
of energy is demanded because the battery is at a low state-of-charge and the vehicle needs
to accelerate. The motor controller also has to be a full ‘four‐quadrant’ type.7 The resulting
hybrid system is illustrated diagrammatically in Figure 12.27. A measurement of the
battery state-of-charge is necessary, rather than just a ‘fully-charged’ indication. The energy
flows to and from the different components of the system that can be quite complex.
Given that the battery in the hybrid system just discussed is supplying fairly short‐term
power peaks and also absorbing power from regenerative braking, the rate of transfer of
charge in and out of the battery is likely to create an operational problem. This is where
the introduction of a ‘supercapacitor’ promises to be particularly beneficial. The specific
7 Four quadrants refer to the four possible modes of motoring, namely, forwards accelerating, braking while
going forwards, backwards accelerating and braking while going backwards.
The Complete System and Its Future
Four-quadrant inverter
Fuel-cell
stack
M
Control
Rechargeable
battery
State-of-charge
measurement
Power
switching
and
control
Control
Power/braking
demand
Figure 12.27 Diagram of a hybrid fuel‐cell–battery system for a vehicle with regenerative braking.
Bold arrows indicate energy flows. A current‐sensing resistor in series with the battery measures the
input and output of charge.
energy (Wh kg−1) of such a device is much less than that of rechargeable batteries, but
the power density is very much greater — typically 2.5 kW kg−1. Supercapacitors can also
sustain at least 500 000 charge–discharge cycles.
The UltraBattery™, invented by the Commonwealth Scientific and Industrial Research
Organisation (CSIRO) in Australia,8 is the first practical example of combing the
attributes of a rechargeable battery — in this case, lead–acid chemistry — and a
supercapacitor into an efficient and affordable energy-storage device. The technology
was primarily designed to meet the high‐rate partial state‐of‐charge duty required from
hybrid electric vehicles (HEVs) but can be reconfigured for a variety of applications, for
example, power tools, forklift trucks, high‐power uninterruptible power supplies,
remote‐area power supplies and grid frequency regulation. The CSIRO technology has
been taken up by the Furukawa Battery Co., Ltd., Japan, and the East Penn Manufacturing
Co., Inc., USA, and is under mass production for both HEV and renewable energy
applications. The UltraBattery™ is under evaluation by many carmakers worldwide and
it has been adopted in both the Honda Odyssey Absolute and the Honda StepWGN
hybrid models as original equipment.
Hybrid electric vehicles have grown in popularity in recent years — driven by improved
fuel utilization and emissions reduction. When fuel cells are employed in hybrid
vehicles, the two basic configurations available are those illustrated in Figure 12.28.
In addition to fuel cell + battery hybrids, there are a large number of other hybrid
options that can include a fuel‐cell stack. An example is that of a solar PV array, fuel cell
and battery. Such systems have been successfully deployed in roadside variable message
signs and could also serve as remote power supplies. The fuel cell may compensate for
the somewhat unreliable nature of solar energy in many situations.
8 Lam, LT, Haigh, NP, Phyland, CG and Rand, DAJ, 2005, High performance energy storage devices,
International Patent WO/2005/027255.
383
384
Fuel Cell Systems Explained
(a)
Battery or
supercapacitor
Ancillary
devices
Regeneration,
recharge battery
when needed
Assist fuel cell when needed
Stored
hydrogen
Fuel-cell
stack
DC/DC
converter
Inverter
Motor
(b)
Ancillary
devices
Stored
hydrogen
Fuel-cell
stack
DC/DC
converter
Battery
Inverter
Motor
Figure 12.28 Diagrams of (a) parallel and (b) series versions of hybrid fuel‐cell vehicle.
12.4
Analysis of Fuel‐Cell Systems
Throughout this book, it has been stressed that many different components are required
to build a complete fuel‐cell system. Whereas the choice of stack technology will be
influenced by the application, the availability of fuel will determine the extent of any fuel
processing that may be required. The application will also dictate the level of power
electronics that is necessary to operate the complete system. Over and above the issues
of system design and construction, the cost of materials and fabrication are keys to the
successful commercialization of fuel‐cell systems. Despite all of their advantages in terms
of environmental impact and performance with respect to other technologies, most fuel‐
cell systems have competitors. Moreover, many of these alternative technologies are
well‐established in mature markets. In the transport sector, for example, it is the internal
combustion engine — a highly engineered technology that is of relatively low cost
compared with fuel‐cell systems. In the stationary power market, large generators based
on gas or steam turbines with lower capital costs can produce electricity more cheaply
The Complete System and Its Future
than most fuel‐cell systems. Such competing technologies set the cost targets that must
be met if fuel‐cell systems are to become commercially viable.
Various analytical methods can be employed to establish the competitiveness of
different designs of fuel cell. Although economic modelling is outside the scope of this
book, the following approaches will be briefly examined as they can help place fuel cells
within the wider context of alternative technologies:
●
●
●
●
Well‐to‐wheels analysis.
Power‐train or drive‐train analysis.
System life‐cycle assessment.
Flowsheet or process modelling.
12.4.1 Well‐to‐Wheels Analysis
The infrastructure required to generate, distribute and store hydrogen for fuel‐cell
vehicles (FCVs) is critical to their commercial success. A number of different routes
for generating hydrogen have been examined in Chapter 10. Several hydrogen supply
options exist for vehicles:
●
●
●
●
●
Hydrogen generated by steam reforming of natural gas in a large centralized plant and
then delivered as liquid hydrogen by trailer to filling stations.
Hydrogen generated by steam reforming of natural gas in a large centralized plant and
then delivered as compressed gas by pipeline to filling stations.
Hydrogen generated as a by‐product, e.g., from oil refineries and industrial ammonia
production plants.
Hydrogen produced at filling stations by small‐scale steam reformers that run on
pipeline natural gas.
Hydrogen produced at filling stations by electrolysers.
The consensus view from several studies in the United States is that, in the near term,
hydrogen generated by steam reforming of natural gas in small, localized reformers is
the best option. Where there is no natural gas supply, hydrogen produced by electrolysis
may be the preferred alternative (especially when using renewable electricity). As a
hydrogen infrastructure emerges in the future, there may come a time when centralized
production would be economic, with the added bonus of being able to collect and
sequester the carbon dioxide that is generated from the reformer. To consider the various
scenarios for road transport applications, well‐to‐wheels (WTW) analyses may be
carried out to quantify the options for generating hydrogen (from the well), its conversion
to electricity on a vehicle and transfer of the power to mechanical energy for driving the
wheels. Consequently, WTW analysis determines the energy efficiency of converting
the energy (in the well) to energy required at the wheel of a vehicle.
In a well‐documented WTW analysis carried out in 2001 by General Motors (GM), the
Argonne National Laboratory (ANL) and others9 for the North American market, 15
different vehicles were investigated. These included conventional vehicles and HEVs
9 Well‐to‐Wheel Energy Use and Greenhouse Gas Emissions of Advanced Fuel/Vehicle Systems,
General Motors, Argonne National Laboratory, BP, Exxon/Mobil and Shell, April 2001. Available from
http://www.ipd.anl.gov/anlpubs/2001/04/39097.pdf (accessed 28 September 2017).
385
386
Fuel Cell Systems Explained
8
Gasoline
Diesel
Naptha
Natural gas
North American
and non-North
American
39
Compressed natural gas
Methanol
Hydrogen
Liquid fuels made using
Fisher-Tropsch reactors
Renewable
Corn
Woody
Herbaceous
12
Ethanol
16
Hydrogen
compressed
and liquid
Crude
oil
Electricity
Combined cycle, hydro
nuclear
North Europe, CA,
and US mixture
Figure 12.29 The 75 different pathways investigated in the North American well‐to‐wheels study
conducted in 2001 by GM, ANL, BP, Exxon/Mobil and Shell. (Source: Reproduced with the permission
of Elsevier.)
with spark‐ignition and compression‐ignition engines, as well as hybridized and non‐
hybridized FCVs with and without on‐board fuel processors. All 15 vehicles were
configured to meet the same performance requirements. Thirteen fuels, selected from
75 different fuelling options or pathways, were considered in detail. These included
low‐sulfur gasoline, low‐sulfur diesel, crude oil‐based naphtha, Fischer–Tropsch (FT)
naphtha,10 liquid or compressed gaseous hydrogen (based on five different pathways for
production), compressed natural gas, methanol and neat and blended (E85) ethanol.
The 75 different pathways for the WTW analyses undertaken in the study, as shown
Figure 12.29.
There are some key findings from the North American study that are worth examining
here, since they may influence the development of infrastructure in many countries that
are seeking to adopt hydrogen vehicles. These findings are as follows:
Total energy use. For the same amount of energy delivered to the vehicle tank,
petroleum‐based fuels and compressed natural gas exhibit the lowest energy losses
from the well-to-tank (WTT). FT naphtha, FT diesel, gaseous hydrogen from
natural gas, methanol and corn‐based ethanol are all subject to moderate WTT
energy losses. By contrast, liquid hydrogen from natural gas, hydrogen from
electrolysis (gaseous and liquid), electricity generation and cellulosic bioethanol
suffer large WTT energy losses.
10 The Fischer–Tropsch process is used for making fuels artificially. Basically a fuel such as biomass, or
even natural gas, is steam reformed using the methods described in Chapter 10. The resulting hydrogen and
carbon dioxide products are then reacted, over catalysts developed by Fischer and Tropsch, to produce
liquid fuels such as octane (C8H18), nonane (C9H20) and decane (C10H22).
The Complete System and Its Future
Greenhouse gas (GHG) emissions. Liquid hydrogen (produced in both central plants and
filling stations) and compressed gaseous hydrogen obtained by electrolysis can both be
energy inefficient and lead to large emissions of GHGs. Ethanol (derived from renewable
cellulose sources such as corn) offers a significant reduction in GHG emissions. Other
fuel options were found to have moderate energy efficiencies and GHG emissions.
Tank‐to‐wheels (TTW) efficiency. Fuel‐cell systems consume less energy than conventional
power-trains, because of the intrinsic high efficiency of the stacks. Fuel‐cell vehicles
operating directly on liquid or compressed gaseous hydrogen exhibit significantly
higher fuel economy than those employing on‐board fuel processors.
Overall well‐to‐wheels efficiency. Hybrid systems offer consistently higher fuel economy
than conventional vehicles.
There has been a plethora of WTW analyses of the North American vehicle fleet in the
years that have followed the earlier investigation conducted by General Motors et al. With
respect to FCVs, there have been no significant changes in the findings. Well‐to‐wheels
studies that have been carried out for European countries, Japan and elsewhere have often
lead to similar conclusions.11 That is, optimum results are realized when renewable energy
sources such as wind, solar or biomass are used in the production of hydrogen. To a lesser
extent, natural gas vehicles offer improvements relative to hybrid vehicles and are less
problematic in engineering terms than conventional engines adopted to run on hydrogen.
In the context of vehicle efficiency, it is noteworthy that the addition of a small amount
of hydrogen can enhance the combustion of liquid fuels in gasoline and diesel engines.
Although not yet promoted among leading vehicle makers, several studies have shown
that engine efficiency can be improved slightly and emissions reduced by injecting
hydrogen with the fuel. Research is underway to generate the hydrogen required for this
treatment by utilizing heat within the exhaust gas to reform some of the liquid fuel that
is carried on board.
12.4.2
Power‐Train Analysis
In a conventional gasoline‐ or diesel‐fuelled vehicle, energy from the fuel is transmitted
from the engine to the wheels by a mechanical power-train. In an FCV, the power-train
is electrical — energy from the fuel is converted into DC electricity (which may be
converted to AC) to power the motor(s). In an HEV, the power-train may involve a
combination of electrical and mechanical energy conversion. A power‐train or drive‐
train analysis is simply quantification of the energy transfer in the vehicle from fuel to
wheels — it is the final stage in the WTW pathway. The analysis can help define the
relative sizes of the components required for specific needs. For example, urban delivery
vehicles may employ a series hybrid power-train (cf. Figure 12.28b) in which the battery
needs to be large as it is providing most of the power for stopping and starting (short
journeys), and therefore the fuel cell can be small. In such a system, the fuel cell is acting
principally as a charger to keep the battery topped up. By contrast, in a parallel hybrid
drive-train (cf. Figure 12.28a), most of the power may be produced by the fuel cell,
which serves also to keep the battery charged for whenever rapid acceleration is
11 Grube, T, Hohlein, B, Stiller, C and Weindorf, W, 2010, Systems analysis and well‐to‐wheels studies,
in Stolten, D (Ed.), Hydrogen and Fuel Cells, Wiley‐VCH Verlag GmbH, Weinheim, pp. 831–852.
387
388
Fuel Cell Systems Explained
required. Thus, the parallel hybrid may be a better option for long‐range vehicles.
In power‐train analysis, other issues may also be considered. For example, whether one
electric motor is to be employed, with the power transmitted mechanically to the wheels
via a conventional axle differential, or whether it is beneficial to power each wheel with
a dedicated hub motor, thereby reducing mechanical energy losses.
12.4.3
Life‐Cycle Assessment
WTW and power‐train analyses are specific examples of what is generally referred to as
a ‘life‐cycle assessment’ (LCA) (also known as ‘life‐cycle analysis’ and ‘cradle‐to‐grave’
analysis). The technique assesses environmental impacts associated with all the stages
of a product’s life from cradle-to-grave, i.e., from raw material extraction to materials
processing, manufacture, distribution, use, repair and maintenance, and disposal or
recycling at the end of life. Life‐cycle assessments are employed widely to:
●
●
Draw up an inventory of relevant energy and material inputs, as well as emissions to
the environment.
Evaluate the potential consequences associated with identified inputs and outputs of
a process or product.
The information obtained from a LCA study is used to improve processes, support
policy and provide a sound basis for informed decisions on, for example, technology
development. Fuel‐cell systems are candidates for LCAs because they impact the environment in terms of inputs (fuel) and outputs (emissions) throughout their working life and
because the materials of construction have a cost associated with extraction, assembly
and disposal at the end of life. For example, in an LCA of a PEMFC system, the combined
cost of the mining and extraction of platinum, which is required for the catalysts, and of
that incurred in the recovery of the metal at the end of the stack lifetime has to be factored
in to the life‐cycle cost of the system. The same consideration has to be given to other
catalysts employed in the fuel‐processing stages of a system. In this respect, the material
in an MCFC, for example, will involve lower costs in extraction and recovery after service
than those required for a PEMFC or a phosphoric acid fuel cell (PAFC). In addition, the
power that can be generated per kg of catalyst may be much more for the MCFC than for
the PEMFC, and therefore the life‐cycle cost of the MCFC may be lower than that of the
PEMFC. Such issues have to be considered in a full LCA of any fuel‐cell system and may
help identify the best pathway for system development.
The procedures undertaken in an LCA are described in ISO 14040 : 2006 and
14044 : 2006, which are included in the ISO 14000 family of Environmental Management
Standards set by the International Organization for Standardization (ISO). The standards
distinguish four phases of an LCA:
1) Goal defining and scope
● Define the purpose of the activity.
● Define boundary conditions.
● Define the life cycle of the product, process or activity.
● Recognize the general material flow in the life cycle.
● Identify all operations that contribute to the life cycle and fall within the system
boundaries, including time and spatial boundaries.
The Complete System and Its Future
2) Inventory
● Quantify energy, raw materials and environmental releases throughout all stages
of the life cycle.
● Define scope and boundaries by gathering data and running computer models.
● Analyse results and draw conclusions.
3) Impact assessment
● Characterize the effects of resource requirements and environmental loadings
identified.
● Address ecological and human health impacts.
4) Improvement assessment
● Evaluate needs and opportunities to reduce environmental burdens throughout
the whole life cycle.
Since a full LCA can be both a costly and time‐consuming exercise, only well‐established
and large organizations with a budget for sustainability studies are capable of completing
such a task. More often, WTW or process studies are undertaken as a means of
comparing one fuel‐cell system with another.
12.4.4
Process Modelling
Fuel‐cell power plants for the stationary market can be divided broadly into two
categories: power‐only systems and cogeneration systems. The former include backup
or uninterruptible power systems running on hydrogen, as well as various niche systems
that may be fuelled with alternative fuels such as ammonia or methanol. Examples of
power‐only systems can be found operating in telecommunication towers and similar
remote applications. By contrast, cogeneration systems produce useful heat (or cooling)
as well as electricity and are therefore more complex. Portable versions of these systems
for consumer electronic items present their own unique issues relating to miniaturization of the fuel processing and close integration of the stack and control system, such
issues have been discussed in Chapters 4–6.
In process modelling, the designer starts with a PFD and carries out an analysis of the
steady‐state operation of the plant. Once this is complete, consideration is given to how
the stack is to be started up and shut down, how it responds to load changes during
operation and what eventuates as the cell degrades. All of these features can be modelled
using computer programmes such as Hysis™, MATLAB®, Simulink®, Cycle‐Tempo
(Technical University of Delft) and Aspentech™. The last‐mentioned programme will
model steady‐state processes as well as dynamic changes. TRNSYS is a useful platform
for dynamic modelling of thermal energy systems and can show how fuel cells will
integrate with other energy generators, e.g., PV cells, wind turbines and electrolysers.
Once the fuel‐cell system has been modelled, the basic requirements for the stack and
each item of the BoP will be understood. Consequently, a specification can be drawn up
for the plant and detailed design can begin.
To illustrate the issues involved in the genesis of a system, the design of a stationary
PEMFC system for application at a scale of around 100 kW is considered in the following
text. The PFD is shown in Figure 12.30. Various assumptions are made as follows:
●
Fuel input — natural gas fed at a flow rate of 1 kmol h−1, i.e., approximately equivalent
to a 100‐kW supply. For simplicity, the system will be modelled in the following for
pure methane.
389
390
Fuel Cell Systems Explained
20
HX-12
Natural
gas
Water
Steam
Reformer
3
2
HX-2
AIR
HX-11
Pre-reformer
18
4
5
7
6
HX-4
HX-3
8
HX-5
HTS
9
12
HX-6
LTS
14
PROX
HX-8
Combustion
Cooling water
19
15
Exhaust
16
Fuel
cell
Fuel cell off-gas 17
Stream
Temp/°C
Pressure/Bar
Mole flow/kmolh
Mass flow/kgh
Enthalpy/kW
Mole flows/kmolh
CH
CO
CO
H O
H
O
N
Nat. Gas Water
Air
2
5
6
7
9
3
25
25
20
300 285.5
750
400
400
150
1.2
1.7
1
1.69
1.66
1.65
1.64
1.62
1.68
1
3 13.2
1 4.021 5.958 5.958 5.958 5.959
16.0
54.0 379.3
16.0
70.1
70.1 70.1
70.1
70.1
–20.92 –239.4
0 –17.43 –210.4 –131.3 –151.1 –155.8 –170.8
1
0
0
0
0
0
0
0
0
0
0
0
0
3
0
0
0
0 2.88
0 10.25
1
0
0
0
0
0
0
0.989
0.011
0
2.989
0.032
0
0
0.021
0.595
0.384
1.637
3.321
0
0
0.021
0.595
0.384
1.637
3.321
0
0
0.021
0.193
0.787
1.234
3.723
0
0
0.021
0.005
0.976
1.045
3.912
0
0
12
15
16
120
80
70
1.61
1.59 1.59
5.959 5.958 11.09
70.1
71.1 318.6
17
18
80
300
1.58
1.57
15.41 15.41
389.8 389.8
19
70
1.6
2.04
58.9
20
775
1.56
17.21
449.6
0.021 0.021
0.005
0
0.976 0.981
1.045
1.1
3.912 3.856
0
0
0
0
0.021
0
0.981
4.379
0.578
0.361
9.09
0
0
0
0
0
0.45
1.59
0
0
1.053
5.1
0
0.38
10.68
0
0
0
0
0
2
9.09
0.021
0
0.981
4.379
0.578
0.361
9.09
Figure 12.30 Process flowsheet for a 100 kW PEMFC system operating in a stationary power
application of about 100 kW output. Note that, for clarity, some stream numbers have been omitted
from the table. HX, heat-exchanger; HTS, high‐temperature shift; LTS, low‐temperature shift.
(Source: Reproduced with the permission of Elsevier.)
●
●
●
●
●
●
●
Steam reforming will be the means of conversion, with an initial steam–carbon ratio
of 3. This is higher than the minimum required thermodynamically to prevent carbon
deposition (see Section 10.4.4, Chapter 10). The reformer operates with an outlet
temperature of 750°C to maximize hydrogen production.
The reactions are all modelled assuming that there are no kinetic limitations,
i.e., reactions reach a state of thermodynamic equilibrium.
Desulfurization of the natural gas is carried out at atmospheric pressure and temperature
by an absorber and is not therefore considered in this example system.
Pre‐reforming is carried out in an adiabatic reactor, operating between 300 and
250°C, to reduce the concentration of hydrocarbons of high molecular weight in the
feed gas to the main reformer reactor.
The feed pressure for natural gas is set at 170 kPa to allow for pressure drops through
each of the fuel‐processor elements.
There are two stages of shift: (i) a high‐temperature reactor with a catalyst of
iron oxide that is operating at an inlet temperature of 400°C and (ii) a low‐temperature
reactor that contains a 30 wt.% CuO, 33 wt.% ZnO and 30 wt.% alumina catalyst
at 200°C.
A preferential oxidation (PROX) unit is modelled in the Aspen flowsheet code by two
stoichiometric reactors, one to perform the PROX of carbon monoxide and the other
to remove the remaining oxygen via reaction with hydrogen. These have been
removed in the flowsheet shown in Figure 12.30, and the PROX reactor is shown as a
single unit, as would be experienced in practice.
The Complete System and Its Future
●
Heat for the reforming reaction is provided by combustion of exhaust gas from the
anode of the fuel cell and is supplemented by fresh natural gas, as required.
A material balance for the flowsheet was obtained by simulating the whole process
using Aspentech software and is summarized in the stream data shown below the PFD
in Figure 12.30. Pinch analysis12 was applied to optimize the layout and connection of
the heat-exchangers. Knowing the flow rates of gas through each of the process units
and the respective catalysts employed, the sizes of the reactors can be calculated and a
preliminary mechanical design formulated. Also once the flow rates and heat loads are
known, the system efficiency can be found from:
Efficiency
AC output power of system
Power of fuel supplieed
Power of fuel (stack efficiency inverter efficiency) paarasitic loads
Power of fuel supplied
The hydrogen is supplied to the stack at a rate of 3856 kmol h−1 (stream 15). This is
equivalent to 259 kW, with respect to the lower heating value (LHV) of 241.83 kJ mol−1.
If it is assumed that the fuel utilization in the stack is 85% and that an operating voltage
of 0.65 V is selected, the efficiency of the stack is found from equation (2.28), Chapter 2
as follows:
Efficiency
f
Vc
1.25
0.85
0.65
1.25
0.442 44.2% (LHV)
(12.22)
Therefore, the electrical output (DC) of the stack is 0.442 × 259 = 114.5 kW. Taking a
representative efficiency of 95% for an inverter, the gross AC power produced by the
stack will be 114.5 × 0.95 = 108.8 kW. Assuming a parasitic power requirement of
5.89 kW for compressors and pumps, the net AC power delivered by the system can
therefore be expected to be 108.8 − 5.89 = 102.9 kW. Given that 1 kg mol h−1 of methane
as supplied to the system has a combustion enthalpy of 802.6 kJ mol−1 (LHV), the net
efficiency of the whole system is therefore:
Efficiency
102.9 3600
802.6 1000
0.462 46.2% (LHV)
(12.23)
It is possible to estimate the size of the stack required as follows. The total current, I,
that has to be delivered by the stack is given by:
I
PDC
Vc
114.5 1000
176 154 A
0.65
(12.24)
where PDC is the electrical DC output of the stack and Vc is the stack operating voltage.
Referring to Figure 3.1, Chapter 3, which gives the expected performance from a good
12 Pinch analysis was introduced in Section 7.2.3.
391
392
Fuel Cell Systems Explained
PEMFC, the current density is expected to be about 600 mA cm−2 if the cell voltage is
0.65 V. The total area of cells that make up the stack will therefore be:
Area
current
current density
176 154
0.6
293 590cm 2
(12.25)
This may seem like a large area, but if the stack is built with cells of a reasonable size,
say, 50 × 50 cm, each cell area would be 2500 cm2, and the total number of cells required
would be 293 590/2500, i.e., approximately 120 cells. In practice, such a system would be
made up of a perhaps two stacks wired in parallel, each containing 60 cells. Limiting
the number of cells in a single stack to around 60 also keeps the stack voltage to a safe
level — in this case, 60 × 0.65 = 39 V.
Once a steady‐state system has been established, the designer of the fuel cell will need
to focus attention on questions such as the following:
●
●
●
●
●
●
Is additional heat required to raise the temperature of some reactors (e.g., the
reformer) up to the operating point on start‐up?
Is a supply of hydrogen required to activate the reforming catalyst before start‐up?
How will the system perform at part‐load or under changing loads? Some dynamic
modelling may be required.
Is a purge gas required when the stack is shutdown or held in a hot standby
condition?
What control elements are required (e.g., control valves, thermocouples and other
sensors)?
What type of control system is needed?
After such issues have been resolved, detailed mechanical and electrical drawings can
be produced. Normally, and especially for large installations, a hazard identification
(HAZID) and/or a hazard and operability (HAZOP)13 study or similar safety risk
assessment and analysis will also be conducted.
12.4.5
Further Modelling
The approach taken in this book has been a pragmatic one. The fuel‐cell systems that
have been described use technologies that have emerged during the course of the
20th century, and chapters have been devoted to particular types of fuel cell. There are,
of course, common features to all fuel‐cell technologies in terms of cell components
(catalyst layers, electrodes, electrolyte) and the chemical and physical reactions that
occur at the interfaces between these components. The processes that occur within a
13 A hazard and operability (HAZOP) study is a standard hazard analysis technique enacted worldwide
by process industries for the preliminary safety assessment of new systems or modifications to existing ones.
The HAZOP study is a detailed examination, by a group of specialists, of components within a system to
determine what would happen if that component were to operate outside its normal design mode.
Each component will have one or more parameters associated with its operation such as pressure, flow rate
or electrical power. The HAZOP study looks at each parameter in turn and imposes guide words to list
possible off‐normal behaviour such as ‘more’, ‘less’, ‘high’, ‘low’, ‘yes’ or ‘no’. The effect of such behaviour is
then assessed.
The Complete System and Its Future
fuel cell can be characterized by using mathematical models. In terms of mass transport,
for example, it is possible to model the following actions: (i) flow velocity within a fuel‐
cell channel, (ii) mass transport in the gas‐diffusion layer (GDL) and catalyst layers and
(iii) proton and water transport in the membrane. Furthermore, heat transfer within the
membrane and electrode layers can also be described.
A one‐dimensional (1D) model is the simplest type of mathematical representation of
a fuel cell. For instance, it ignores losses due to reactant transport in the feed channels
and is a relatively easy to derive using an Excel spreadsheet given the Nernst equation
and the theory discussed in Chapters 2 and 3. A simple 1D model will predict quite
accurately the open‐circuit voltage of cells running on hydrogen, on the assumption
that the cell is isothermal. A two‐ or three‐dimensional model is a more complex
proposition on account of the voltage losses that occur during operation a fuel cell,
as discussed in Chapter 3. Two approaches can be taken in the development of a model
that will account for these losses:
1) Simply use data that has been obtained experimentally to devise an empirical relationship
between voltage and current (as a function of the other operating parameters such as
temperature and pressure). Although widely adopted in the literature, this procedure
does not help in understanding the processes that are occurring within the fuel cell.
2) Apply the Butler–Volmer (or Tafel) equation and knowledge of the materials’
properties to estimate the total cell voltage loss by summing all of the individual
losses at the electrodes and through the electrolyte.
Approach 2 has been adopted with some success for PEMFCs, but a particular difficulty
arises in trying to account for the transport of water.14 With high‐temperature fuel cells,
there is no water‐management issue, and the anode overpotential can be virtually ignored.
The voltage of an SOFC running on hydrogen, for example, can be determined reasonably
accurately from equation (3.10), Chapter 3. Modelling of internal reforming SOFCs (or
MCFCs) is a much more difficult proposition. The mechanism of internal reforming is
subject to much debate, and there are conflicting models in the literature. A simple
approach is to assume that the steam reforming reaction is fast and comes to equilibrium
over the anode (or internal reforming catalyst in the case of the MCFC) before the
occurrence of any electrochemical reactions. Equation (3.10) can then be invoked to
calculate the cell voltage that arises from the reformed equilibrium gas mixture over the
whole of the anode surface. Unfortunately, this approach is rather naive because in a
functional internal reforming cell, the fuel gas composition changes along the channels
from anode inlet to outlet. Water is produced by the anode electrochemical reaction as
hydrogen is consumed, and both these processes affect the reforming reaction. It can be
expected, therefore, that the concentration of hydrogen at the surface of the anode will
decrease on moving from the inlet to the outlet of the cell. While it is possible to measure
the composition of the anode exhaust gas, it is virtually impossible to measure the
composition of the anode gas along the anode channels of an SOFC or a MCFC.
Certainly there is much more to learn about the fundamental processes at the heart of
the fuel cell.
14 Berning, T., Lu, D.M. and Djilali, N., 2002, Three‐dimensional computational analysis of transport
phenomena in a PEM fuel cell, Journal of Power Sources, vol. 106, pp. 284–294.
393
394
Fuel Cell Systems Explained
12.5
Commercial Reality
12.5.1
Back to Basics
In her book, Fuel Cells — Current Technology Challenges and Future Research Needs
(Elsevier, 2012), Noriko Behling has argued that:
‘The difficulty of the technology is rooted in the complexities of how fuel cells
work, which involve multiple chemical and physical interactions at the atomic
level. Perhaps no advanced technology on the market today including airplanes,
computers, or even nuclear reactors require the scale, magnitude, and range
of scientific, physical, and engineering knowledge that fuel cell technology
requires.’
Why is it that fuel cells are always a ‘few years away’ from commercialization,
and why, despite the injection of massive amounts of funding for research and
development, do they never seem to quite make it past the demonstration
stage — except in a few high‐value niche markets? The plain fact is that fuel‐cell
systems are complex and therefore require the input of expertise from many disciplines
in science, engineering and technology. In particular, it has often proved difficult to
bring together the necessary skills to focus on the fundamental issues that could lead
to the required reduction in costs. The reasons for this situation are many; for
example, there has often been a lack of sufficient funding sustained over long periods
and an impatience on the part of developers to launch a product to market, even
when there is no apparent market pull.
Collectively, the various chapters of this book have assessed the status of various
types of fuel‐cell system in terms of their design, components, mode of operation and
performance. Although much is understood by developers in terms of system engineering,
there is still considerably more to learn about the processes taking place in the individual
fuel cell at the atomic and molecular level. What, for example, are the physical and
chemical principles that control the exchange‐current density of a catalyst material, and
why is platinum such an active catalyst for hydrogen oxidation compared with other
metals? In the early days of fuel cells, in the 19th century, electrochemistry was just
emerging, and the physics and chemistry known at the time was unable to address such
questions. In more recent years, with the advent of powerful techniques such as X‐ray
photoelectron spectroscopy (XPS), which is able to identify surface species on catalyst
materials, a greatly enhanced understanding of the processes occurring within fuel cells
has been achieved. For example, XPS has been applied to elucidate how non‐platinum
group catalysts function for PEMFCs. Another tool available to the scientist in the
21st century that did not exist beforehand is the computing power necessary to carry
out complex mathematical calculations. For instance, this facility enables chemical
and physical processes to be modelled at the atomic and molecular level through the
application of quantum mechanical methods such as density functional theory (DFT).
Without such detailed basic research, it is feared that no matter how much money and
time may be poured into product development, progress will be haphazard and ultimate
success will be doubtful. Consequently, practical and affordable fuel cells will remain
always ‘a few years away’.
The Complete System and Its Future
12.5.2
Commercial Progress
The attributes of fuel‐cell systems — high efficiency, low emissions, silent operation — that
were touted as unique features by their pioneering proponents did not prove sufficient
to establish the technology as an alternative to other forms of power generation.
A breakthrough came with the PAFC in the 1990s when another beneficial property
became apparent, namely, that of reliability. With banks and other financial institutions
requiring ‘5 nines’ reliability, i.e., 99.999%, to avoid costly power outages, a PAFC
(or preferably two running in parallel to provide some redundancy) could easily achieve
this target. The high capital cost of the PAFC systems was small in comparison with
the losses that would arise from even a few seconds of power outage for the financial
institutions. A business case for installing systems could therefore be made. More recently,
a PEMFC with stored hydrogen has been shown to operate in sub‐zero temperatures,
and thereby a compelling case can be made for forklift trucks to be powered by fuel cells
in refrigerated warehouses. Both of these high‐quality power and material‐handling
applications are niche markets where the fuel cell has gained some commercial status. In
many other cases, notably road vehicles, similar success has yet to be achieved.
In 1998, Daimler‐Benz announced that it would be producing between 40 000 and
100 000 FCVs by 2004.15 This was shortly after commercialization agreements had been
reached with Ballard Power Systems and other parties. At that time, Ballard was concentrating on demonstrating their fuel‐cell systems in buses. Progress has been much slower
than expected — in terms of public transport, at the end of 2015 there were globally
less than 200 fuel‐cell buses operating and less than 50 minibuses. At the same time,
fewer than 3000 fuel‐cell cars were on the roads worldwide, with OEMs promoting the
cars on a city‐by‐city basis in Japan, the United States (notably California) and the EU,
rather than by country.16 A good case may be made for increased production of fuel‐cell
buses, and if progress continues at the present rate, there are encouraging prospects that
the cost could be brought down to a competitive position. Vehicle manufacturers recognize that, important though they are, battery electric vehicles will not address all the
needs of the private motorist. Accordingly, there is renewed commitment to fuel‐cell cars
by most of the major automakers and growing optimism that such vehicles will prove
cost‐effective once fuelling issues have been addressed. A case can also be made for fuel‐
cell systems in rail transport. For instance, in May 2015, Hydrogenics Corporation signed
a 10‐year agreement to supply PEMFC systems to Alstom Transport, the France‐based
train manufacturer. This has led to the unveiling of the Alstom Coradia iLINT, a hydrogen‐
fuelled multiple unit, in September 2016; see Figure 12.31. As of January 2018, Alstom is
building 14 trains for deployment from December 2021 in Lower Saxony.
The long time that fuel cells are taking to transform into commercial products should
be put into perspective by making a comparison with other game‐changing or disruptive
technologies. Perhaps one of the most informative examples is the evolution of PV cells.
Interestingly, it was in 1839 — the same year Grove assembled the first fuel cell — that
a 19‐year‐old Frenchman, Edmond Becquerel, found that electricity could be produced
directly from sunlight. He measured a voltage and could draw a current between two
platinum electrodes immersed in an acidic solution of silver chloride when the solution
15 All, J., 1998, Auto makers race to sell cars powered by fuel cells, Wall Street Journal, 15 March 1998.
16 This information was provided by Kerry‐Ann Adamson of 4th Energy Wave, 2015.
395
396
Fuel Cell Systems Explained
Fuel-cell composition
Traction motor
Traction inverter and
DC/DC-converter
Auxiliary converter
Hydrogen fuel tank
Battery composition
Figure 12.31 The Alstom Coradia iLint zero‐emission fuel‐cell train.
was illuminated by sunlight. This discovery was followed in 1873 by the English engineer
Willoughby Smith who showed that selenium possesses photoconductivity. A decade
later, the first selenium photocell was produced by the American inventor Charles Fritts,
and it continued to be used as a light sensor until the 1960s. Meanwhile, the advent of a
more practical PV cell awaited the research conducted by Gordon Teal and John Little at
the Bell Telephone Laboratories (nickname ‘Bell Labs’), USA, in the early1950s.17 The
two chemists were the first to grow single crystals of germanium and, later, silicon. Their
work heralded the world of transistors and other semiconductors.
Bell Labs exhibited the first high‐power silicon PV cell in 1954. The first satellite to
use solar power, the US Vanguard 1, was launched in 1958; it employed a 100‐cm2 solar
panel that delivered 0.1 W. Since then, solar cell technology has progressed in terms
of enhanced efficiency, greater manufacturing ability, lower failure rate, improved
lifetime and, most importantly, reduced capital cost. The Japanese company, Kyocera,
was the first manufacturer in the world to mass-produce polycrystalline silicon solar
cells, which were made via a casting method that is today’s industry standard. With
commercial solar systems now achieving significant penetration in the electricity
market, it is easy to overlook the long gestation period for the technology and to forget
that the first niche market, in around 1978, was to power pocket calculators. It is also
important to realize that, as with the different types of fuel cell, there are also several
different types of silicon PV cell, e.g., crystalline, polycrystalline and amorphous. Each
17 Teal, GK and Little, JB, 1950, Growth of germanium single crystals, Physical Review, vol. 78, p. 647.
The Complete System and Its Future
version of silicon solar cell has had its own line of development and the progressive
advancement of one technology has influenced another. Studies since 1990 have shown
that the impressive cost‐reduction of silicon cells from US$76 per watt in 1977 to
US$0.2 per watt in 2017 has taken place in a number of steps, as the science and manufacturing capability improved. Perhaps there is a lesson here for developers of fuel cells.
12.6
Future Prospects: The Crystal Ball Remains Cloudy
A short history of the fuel cell in the 19th and early 20th centuries has been given in
the opening chapter of this book. Much of the development of the technology took
place in Europe and later in the United States. The work culminated with the fuel cells
developed for the US space programme of the 1960s. United Technologies Corporation
(UTC) was set up by the US‐based firm of Pratt and Whitney in 1958 to develop the
alkaline fuel cell pioneered in the United Kingdom by Francis Bacon. In 1966, the
company supplied the fuel cells to the National Aeronautics and Space Administration
(NASA) for the Apollo project and later for the Space Shuttle missions until 2010.
Having developed PAFCs during the 1970s and early 1980s, the UTC fuel‐cell team
became focused on the technology business and in 1985 formed a wholly owned
trading subsidiary under the name International Fuel Cells (IFC). The enterprise was
subsequently renamed UTC Fuel Cells and finally UTC Power in 2001. The activities
of UTC fuel cells were targeted towards stationary systems and resulted in the
marketing of the PC25 200‐kW and later 400‐kW packaged systems. The company
later expanded its interests to include PEMFCs for transport. Thus UTC Power could
trace its history back to the pioneering work of Bacon, but by the early years of the 21st
century, it was clear that the PAFCs produced by the company were not economically
viable for most applications for which they were designed (i.e., cogeneration). Only in
some niche markets (e.g., quality power for data centres and banks) could a robust
business case be made.
A brand new venture, Quantum Leap Technology, was formed in 2003 by the
innovator Brett Vinsant who had developed a PEMFC in his garage in Hillsboro, Oregon.
In August 2005, Quantum Leap changed its name to ClearEdge Power and directed its
efforts to the development of novel PEMFC systems that were built using semiconductor
technology. The company went through several rounds of seed and venture funding, so
in May 2007, it had grown to having 20 employees and had raised US$10 million in
venture capital. In early 2008, ClearEdge sold and installed its first fuel‐cell system and
in January 2009 raised another US$11 million in venture capital. Doubling its workforce
in less than 12 months, further successful funding rounds continued in 2010 with orders
for 5‐kW backup power plants. In June 2010, ClearEdge signed a US$40 million deal to
supply 800 fuel‐cell systems over a 3‐year period to LS Industrial Systems, a subsidiary
of the South Korean LS Group. To undertake and complete this assignment, even more
venture capital was raised (over US$70 million).
ClearEdge presented itself as a rapidly growing clean technology start‐up business.
Several other examples can be found in the fuel‐cell industry, some more successful
than others. Companies cannot, however, continue to be supported by venture capital
indefinitely. Hungry for expansion, ClearEdge acquired some of the assets of UTC
Power in February 2013. In all such deals, there are winners and losers, and in this case
397
398
Fuel Cell Systems Explained
it was UTC Power who ended up filing for bankruptcy protection in 2014. This story
ended in July 2014 when ClearEdge was purchased by the Doosan Group, a South
Korean conglomerate.
The purpose of relating the fate of UTC Power and Clear Edge Power is not to highlight particular companies, but to show that the fuel‐cell industry is far from mature. It
is a fragile industry because few of the companies have firm order books and the infrastructure to deliver mass‐produced products. A small number of the companies that are
currently trading have focused on fuel‐cell systems for more than 25 years, e.g., Rolls‐
Royce Fuel Cells, Fuel Cell Energy and Ballard Power Systems. Many more companies
are very much younger and have been spun out of research groups who have claimed a
particular breakthrough, or they have been set up by entrepreneurs seeking to exploit
the technology to fulfil a new business model in a carbon‐constrained world. Fuel‐cell
technology — even the advanced PAFC and PEMFC systems — finds it difficult to
compete in the market place with alternative entrenched power-generation technologies
(e.g., gas micro‐turbines) despite some apparent advantages. The reader can explore the
fate of other fuel‐cell companies, such as the various SOFC developers (e.g., Siemens,
Westinghouse, Sulzer Hexis) and MCFC developers (MC‐Power and MTU‐Onsite),
who have fallen by the wayside.
For many years, particularly in the closing years of the 20th century, fuel‐cell
systems were oversold by zealous engineers. Hype led to the prospect of commercial
systems and unfulfilled expectations led to disillusionment by government and other
funding agencies, and to an extent, the public at large. Companies and research
organizations that remain active in the field continue to battle against competing
technologies (e.g., the fuel‐cell car vs. the battery car,18 or stationary natural gas fuel‐
cell systems vs. gas turbines). At the same time, there is a realization that fuel cells will
only achieve widespread commercial success as the understanding of the science and
technology grows.
When the first edition of this book was published (2000), the expected lifetime of a
PEMFC stack was around 2000 h. As the mechanisms of cell degradation have become
more understood so has the proven lifetime increased – it can now be in excess of
20 000 or even 100 000 h in stationary energy-storage applications. At the same time,
improved materials and manufacturing methods are helping to reduce costs.
Consequently, the automotive industry has been encouraged to introduce FCVs such
as the Nexo, which has an impressive range of 600 km and was launched by Hyundai
in early January 2018. We hope, therefore, that this third edition will provide a timely
account of the basic science and technology of fuel-cell systems, together with an
understanding of their challenges and prospects.
18 There is a somewhat spurious argument between proponents of battery electric vehicles and fuel‐cell
vehicles. Both are electric vehicles, have much in common and are anticipated to satisfy different market
sectors. Battery‐only electric vehicles are likely to be commercially viable for short journeys given the
required weight of batteries and limitations in the rate of charging. Hydrogen fuel‐cell vehicles, by contrast,
can be charged quickly and can be expected to have a longer range.
The Complete System and Its Future
Further Reading
Bagotsky, VS, 2012, Fuel Cells: Problems and Solutions, 2nd ed., John Wiley & Sons, Inc.,
Hoboken, NJ.
Behling, N, 2012, Fuel Cells: Current Technology and Future Research Needs, Elsevier,
Burlington, MA.
Dell, RM and Rand, DAJ, 2004, Clean Energy, The Royal Society of Chemistry, Cambridge.
Dell, RM, Moseley, PT and Rand, DAJ, 2014, Towards Sustainable Road Transport,
Elsevier, Amsterdam.
Evers, AA, 2010, The Hydrogen Society…More Than Just a Vision? Hydrogeit Verlag,
Oberkraemer.
Kulikovsky, AA, 2010, Analytical Modelling of Fuel Cells, Elsevier BV, Amsterdam.
Rand, DAJ and Dell, RM, 2008, Hydrogen Energy: Challenges and Prospects, The Royal
Society of Chemistry, Cambridge.
Romm, JJ, 2005, The Hype about Hydrogen: Fact and Fiction in the Race to Save the
Climate, Island Press, Washington, DC.
Schewe, PF, 2007, The Grid: A Journey Through the Heart of Our Electrified World, Joseph
Henry Press, Washington, DC.
Scott, DS, 2007, Smelling Land: The Hydrogen Defense Against Climate Catastrophe,
Canadian Hydrogen Association, Vancouver, BC.
Sigfusson, TI, 2008, Planet Hydrogen — The Taming of the Proton, Coxmoor Publishing
Company, Oxford.
Sperling, D and Cannon, JS, 1994, The Hydrogen Energy Transition: Cutting Carbon from
Transportation, Elsevier Academic Press Pt Inc., San Diego, CA.
Topler, J and Jochen, L, 2016, Hydrogen and Fuel Cell: Technologies and Market
Perspectives, Springer, Berlin. ISBN: 978‐3662449714.
Weiss, MA, Heywood, JB, Drake, EM, Schafer, A and Yeung, AFF, 2000, On the Road on
2020 A Life‐Cycle Analysis of New Automobile Technologies, Energy Laboratory
Report # MIT EL 00‐003, Energy Laboratory, Massachusetts Institute of Technology,
Cambridge, MA, 02139–04307.
399
401
Appendix 1
Calculations of the Change in Molar Gibbs Free Energy
A1.1
Hydrogen Fuel Cell
This appendix shows how to calculate the change in molar Gibbs free energy for the
reaction:
H2
1
2 O2
H2O
(A1.1)
The Gibbs free energy (G) of a system (also known as Gibbs energy or Gibbs function)
is defined in terms of the enthalpy (H), temperature (T ) and entropy (S) according to the
relationship:
G
(A1.2)
H TS
Similarly, the molar Gibbs free energy of formation ( g ), the molar enthalpy of formation
(h f ) and the molar entropy (s)1 are connected by the equation:
f
gf
hf
(A1.3)
Ts
In the case of the hydrogen oxidation reaction (A1.1), it is the change in energy that is
important, i.e., the difference in energy between the reactants (hydrogen and oxygen)
and the product (water or steam). Also, in a fuel cell, the temperature can be taken as
constant2 and, therefore, the following holds:
gf
hf
(A1.4)
T s
The value of ∆h f is the difference between h f of the products and h f of the reactants.
Thus, for the hydrogen oxidation reaction:
hf
hf
H2 O
hf
1
H2
2
hf
O2
(A1.5)
1 Because entropy can be measured as an absolute value, that is, not relative to those of the elements in
their reference states, there is no need to use the term ‘entropy of formation’, but simply use the absolute
entropies for products and reactants.
2 The heat generated by a fuel cell would give rise to an increase in temperature were it not for the fact that
the cell is cooled by the cathode air, by virtue of conducting internal reforming or by a combination of
methods. Enough cooling must be applied to ensure that there are no large temperature gradients that could
cause stresses and therefore degradation of cell materials. Given these requirements, to a first
approximation, the cell may be considered to be at constant temperature.
Fuel Cell Systems Explained, Third Edition. Andrew L. Dicks and David A. J. Rand.
© 2018 John Wiley & Sons Ltd. Published 2018 by John Wiley & Sons Ltd.
402
Appendix 1: Calculations of the Change in Molar Gibbs Free Energy
Similarly, Δs is the difference between s of the products and s of the reactants.
Consequently, for the reaction under consideration:
s
s
H2 O
s
1
H2
2
s
(A1.6)
O2
The values of h f and s vary with temperature according to equations (A1.7) and (A1.8)
given in the following text. These standard equations are derived using thermodynamic
theory, and their proof can be found in textbooks on engineering thermodynamics.3
The subscript to h and s is the temperature, and cP is the molar heat capacity at constant
pressure. Standard temperature is taken as 298.15 K.
The molar enthalpy of formation at temperature T is given by
T
hT
h298.15
(A1.7)
cP dT
298.15
The molar entropy is given by:
T
sT
s298.15
1
cP dT
T
298.15
(A1.8)
The values for both the molar enthalpy of formation and the molar entropy of formation
at 298.15 K are obtainable from thermodynamics tables. Typical data at standard pressure
are given in Table A1.1.4
The molar heat capacity varies with temperature. To use equations (A1.7) and (A1.8),
therefore, it is necessary to know the c p values at constant pressure over a range of
temperatures. Fortunately, empirical equations for the dependence of c p on temperature
are available in many thermodynamics texts.5 The results given by the following three
equations are accurate to within 0.6% over the range 300–3500 K.
Table A1.1 Values of hf are in J mol−1 and those for s in J mol−1 K−1,
both at 298.15 K, for the hydrogen fuel‐cell reaction (A1.1).
hf
s
H2O (liquid)
−285 838
70.05
H2O (steam)
−241 827
188.83
H2
0
130.59
O2
0
205.14
3 For example: Balmer, RT, 2011, Modern Engineering Thermodynamics, Academic Press, New York;
Smith, JM, Van Ness, HC and Abbott, MN, 2005, Introduction to Chemical Engineering Thermodynamics,
7th edition, McGraw Hill Higher Education, Boston, MA.
4 From: Keenan, JH and Kaye, J, 1948, Gas Tables, John Wiley & Sons, Inc., New York.
5 For example: Van Wylen, GJ and Sonntag, RE, 1986, Fundamentals of Classical Thermodynamics,
3rd edition, John Wiley & Sons, Inc., New York, p. 688.
Appendix 1: Calculations of the Change in Molar Gibbs Free Energy
Table A1.2 Sample values for Δhf and Δgf in J mol−1, and Δs in J mol−1 K−1,
for the reaction (A1.2).
Temperature
Δhf
Δgf
Δs
100
−242.6
−0.0466
−225.2
300
−244.5
−0.0507
−215.4
500
−246.2
−0.0533
−205.0
700
−247.6
−0.0549
−194.2
900
−248.8
−0.0561
−183.1
Temperatures are in celsius.
For steam:
143.05 58.040T 0.25 8.2751T 0.5 0.036989T
cP
(A1.9)
For hydrogen:
56.505 22 222.6T
cP
0.75
116 500T
1
560 700T
1.5
(A1.10)
For oxygen:
37.432 2.0102 10 5T 1.5 178 570T
cP
1.5
2 368 800T
2
(A1.11)
All of the equations for cP are in J mol−1 K−1. The values can be substituted into equations
(A1.7) and (A1.8) to yield functions that can be readily integrated and thus evaluated at
any temperature T. This mathematics is undertaken to derive values for Δh and Δs for
steam, hydrogen and oxygen. The values are then substituted into equations (A1.5) and
(A1.6) to give values for ∆h f and Δs that are finally substituted into equation (A1.4) to
calculate the change in molar Gibbs energy of formation, ∆g f . Sample values are given
in Table A1.2.
For liquid water, the standard values for h f and s at 25°C are taken from Table A1.1.
To find h f s and s at 80°C, equations (A1.7) and (A1.8) are again employed, but since the
temperature range (25–80°C) is small, it can be assumed that cP is constant.
A1.2
Carbon Monoxide Fuel Cell
It is possible that in the high‐temperature fuel cells introduced in Chapter 6, the carbon
monoxide generated from steam reforming of a fuel (e.g., methane) is directly oxidized.
The reaction is:
CO
1
2 O2
CO2
(A1.12)
The method used, and the theory employed, for calculating the change in molar Gibbs
free energy, is exactly the same as for the hydrogen fuel cell, except that the equations are
403
404
Appendix 1: Calculations of the Change in Molar Gibbs Free Energy
Table A1.3 Values of hf in J mol−1 and s
in J mol−1 K−1, both at 298.15 K, for the carbon
monoxide fuel‐cell reaction (A1.12).
hf
O2
s
0
205.14
CO
−110 529
197.65
CO2
−393 522
213.80
altered to fit the new reaction. Oxygen features as shown in equation (A1.11), and the
values of the molar heat capacity for carbon monoxide and carbon dioxide are given by:
For carbon monoxide:
cP
69.145 0.022282T 0.75 2007.7T
0.5
5589.64T
0.75
(A1.13)
For carbon dioxide:
cP
3.7357 3.0529T 0.5 0.041034T 2.4198 10 6 T 2
(A1.14)
Together with values from Table A1.3, these equations are used with equations (A1.7)
and (A1.8) to determine the molar enthalpies and molar entropies for the three gases;
see Table A1.3.
The change in the molar enthalpy and the molar entropy is then determined, respectively, from the following two equations:
hf
s
hf
s
hf
CO2
CO2
s
1
CO
1
CO
2
s
2
hf
(A1.15)
O2
(A1.16)
O2
The change in molar Gibbs free energy of formation is then calculated, as for the hydrogen
fuel cell, via equation (A1.4). Some example results are given below in Table A1.4.
Table A1.4 Sample values for Δhf and Δgf in J mol−1, and Δs
in J mol−1 K−1, for the carbon monoxide fuel‐cell reaction (A1.12).
Temperature
Δhf
Δgf
Δs
100
−283.4
−250.7
−0.0877
300
−283.7
−232.7
−0.0888
500
−283.4
−214.6
−0.0890
700
−281.8
−196.5
−0.0877
900
−281.0
−178.5
−0.0822
Temperatures are in celsius.
405
Appendix 2
Useful Fuel‐Cell Equations
A2.1
Introduction
This appendix presents the derivation of useful equations that relate to the following
fuel‐cell parameters:
●
●
●
●
●
●
Oxygen and air usage rate.
Inlet air flow rate.
Exit air flow rate.
Hydrogen usage and energy content of hydrogen.
Rate of water production.
Heat production.
The term ‘stoichiometric’ is encountered in the ensuing discussion. Its meaning could
be defined as ‘just the right amount’. For instance, in the simple fuel‐cell reaction:
H2
1
2 O2
H2O
(A2.1)
exactly two moles of hydrogen would be provided for each mole of oxygen. These
reagents would produce exactly 4F of charge since 2 moles of electrons are transferred
for each mole of hydrogen. Either hydrogen, oxygen or both are often supplied at greater
than the stoichiometric rate. This is especially the case with oxygen if it is being supplied
as air. Otherwise, the air leaving the cell would be completely devoid of oxygen. Note
also that reactants cannot be supplied at less than the stoichiometric rate.
The stoichiometry can be expressed as a variable that is normally denoted by the
symbol λ. Thus, if the rate of usage of a chemical in a reaction is ṅ moles per second,
then the rate of supply is λṅ moles per second.
To increase the usefulness of the equations derived in the following, they have been
formulated in terms of the electrical power of the whole fuel‐cell stack, Pe, and the average
voltage of each cell in the stack, Vc. The electrical power will nearly always be known
because it is the most basic and important information about a fuel‐cell system. If Vc is
not given, it can be assumed to be between 0.6 and 0.7 V — most fuel cells operate in
this region (see Figures 3.1 and 3.2, Chapter 3). The value of Vc could be calculated by
using equation (2.5), Chapter 2, provided the efficiency is known; otherwise, taking Vc
as 0.65 V would be a good approximation. If, however, the fuel cell is pressurized, then
a somewhat higher estimate of Vc should be taken.
Fuel Cell Systems Explained, Third Edition. Andrew L. Dicks and David A. J. Rand.
© 2018 John Wiley & Sons Ltd. Published 2018 by John Wiley & Sons Ltd.
406
Appendix 2: Useful Fuel‐Cell Equations
A2.2
Oxygen and Air Usage
In the basic operation of the hydrogen fuel cell, four electrons are transferred for each
mole of oxygen; see equation (1.3), Chapter 1. Hence, for a single cell:
Charge transferred 4 F amount of O2
(A2.2)
Dividing by time and the rearranging gives:
Oxygen usage
I
mol s
4F
1
(A2.3)
where I represents the current.
For a stack of n cells:
Oxygen usage
In
mol s
4F
1
(A2.4)
It would be more useful, however, to have the formula in kg s−1 so that it is not necessary to know the number of cells, as well as in terms of power rather than current. If the
voltage of each cell in the stack is Vc, then:
Power, Pe Vc I n
(A2.5)
Thus, the current is given by
I
Pe
Vc n
(A2.6)
Substituting this expression into equation (A2.4) gives
Oxygen usage
Pe
mol s
4Vc F
1
(A2.7)
Changing from mol s−1 to kg s−1
Oxygen usage
32 10 3 Pe
4 Vc F
8.29 10
Vc
8
Pe
kg s
1
(A2.8)
This formula allows determination of oxygen usage by any fuel‐cell system of given
power. When Vc is not known, it can be calculated from the efficiency, and if this parameter
is not given, as noted earlier, 0.65 V can be used as a good approximation.
The oxygen will normally be supplied in the form of air, and therefore it is necessary
to adapt equation (A2.7) to air usage. The molar proportion of air that is oxygen is 0.21,
and the molar mass of air is 28.97 × 10−3 kg mol−1. Consequently, equation (A2.7)
becomes:
Air usage
28.97 10 3 Pe
0.21 4 Vc F
3.58 10
Vc
7
Pe
kg s
1
(A2.9)
If the air was used at this rate, then as it left the cell it would be completely devoid of
any oxygen. This is impractical and consequently the airflow is set at a value well above
Appendix 2: Useful Fuel‐Cell Equations
the stoichiometric requirement — typically, twice as much. If the stoichiometry is λ,
equation (2.9) becomes
Air usage
3.58 10 7
Vc
Pe
kg s
1
(A2.10)
The kilogram per second is not, in fact, a very commonly used unit of mass
flow. The following conversions of the mass flow unit to volume under standard
conditions will be found more useful. The mass flow rate from equation (A2.10)
should be multiplied by:
●
●
●
●
3050 to give flow rate in standard m3 h−1.
1795 to give flow rate in standard ft3 min−1, abbreviated as SCFM (i.e., standard cubic
foot per minute).
5.1 × 104 to give flow rate in standard L min−1.
847 to give flow rate in standard L s−1.
A2.3
Exit Air Flow Rate
It is sometimes important to distinguish between the inlet flow rate of the air, which is
given by equation (A2.10), and the outlet flow rate. This is particularly important when
calculating the humidity, which is an issue in certain types of fuel cell — especially
proton‐exchange membrane fuel cells (PEMFCs). The difference is caused by the consumption of oxygen. There will usually be more water vapour in the exit air, but ‘dry air’
is being considered at this stage of the discussion. Water production is examined later
in Section A2.5.
Clearly:
Exit air flow rate inlet air flow rate oxygen usage
(A2.11)
Using equations (A2.10) and (A2.8), equation (A2.11) becomes
Exit air flow rate 3.5 10
A2.4
7
Pe
Vc
8.29 10
8
Pe
kg s
Vc
1
(A2.12)
Hydrogen Usage
The rate of usage of hydrogen is derived in a way similar to that for oxygen, except that
there are two electrons from each mole of hydrogen. Equations (A2.4) and (A2.7) thus
become, respectively:
H2 usage
In
mol s
2F
H2 usage
Pe
mol s
2Vc F
1
(A2.13)
1
(A2.14)
407
408
Appendix 2: Useful Fuel‐Cell Equations
Table A2.1 Energy content of hydrogen fuel expressed in different forms.
Form
Energy content
Specific enthalpy (HHV)
1.43 × 108 J kg−1
Specific enthalpy (HHV)
39.7 kWh kg−1
Effective specific electrical energy
26.8 × Vc kWh kg−1
Energy density as STP (HHV)
3.20 kWh m–3 = 3.20 Wh SL−1
Energy density as STP (HHV)
3.29 kWh m–3 = 3.29 Wh SL−1
SL, standard litre.
To obtain the lower heating value (LHV), multiply the HHV by 0.846.
The molar mass of hydrogen is 2.02 × 10−3 kg mol−1, so equation (A2.14) becomes:
H2 usage
2.02 10 3 Pe
2Vc F
1.05 10
8
Pe
kg s
Vc
1
(A2.15)
under stoichiometric conditions. Obviously, this formula only applies to a hydrogen‐fed
fuel cell. In the case of a mixture of hydrogen and carbon monoxide derived from a
reformed hydrocarbon, the situation will be different and dependent on the proportion
of carbon monoxide present. The result can be transformed to a volume rate by using
the density of hydrogen, which is 0.084 kg m−3 at normal temperature and pressure
(NTP, 293.15 K and 1 atm).
In addition to the rate of consumption of hydrogen, it is often also important to know
the electrical energy that could be produced from a given mass or volume of hydrogen.
The list in Table A2.1 gives the energy in kilowatt‐hours (kWh), rather than in Joules, as
this measure is commonly used for electrical power systems. In addition to the ‘raw’
energy per kilogram and standard litre, there is an ‘effective’ energy that takes into
account the efficiency of the cell and is expressed in terms of Vc, the mean voltage of
each cell. If efficiency of a hydrogen fuel cell has to be taken into account, then the
formula for efficiency derived in Section 2.5, Chapter 2, can be used, namely:
Efficiency
Vc
1.48
(A2.16)
Note that in equation (A2.16), the term for fuel utilization is not included as most pure
hydrogen fuel cells will be assumed to run with 100% fuel utilization.
A2.5
Rate of Water Production
In a hydrogen‐fed fuel cell, water is produced at the rate of one mole for every two
electrons (see Section 1.1, Chapter 1) and can be expressed by adapting equation (A2.7)
to obtain:
Water production
Pe
mol s
2Vc F
1
(A2.17)
Appendix 2: Useful Fuel‐Cell Equations
The molecular mass of water is 18.02 × 10−3 kg mol−1; therefore:
Water production
9.34 10
Vc
8
Pe
kg s
1
(A2.18)
In a hydrogen‐fed fuel cell, the rate of water production is approximately stoichiometric. If, however, the fuel is a mixture of carbon monoxide with hydrogen, then
the water production would be less, namely, in proportion to the amount of carbon
monoxide present in the mixture. For a hydrocarbon fuel that was internally
reformed, some of the product water would be used in the reformation process. For
instance, it was shown in Chapter 9 that if methane is internally reformed, then half
the product water is used in the reforming process, thus halving the exit rate of water
from the fuel cell.
By way of an example, consider a 1‐kW fuel cell that operates for 1 h at a cell voltage
of 0.7 V. This performance corresponds to an efficiency of 47% (with respect to the
HHV), as given by equation (A2.16). Substituting this value into equation (A2.17)
gives:
Water production
9.34 10 8 1000
1.33 10 4 kg s
0.7
1
(A2.19)
It follows that the mass of water produced in 1 h is 1.33 × 10−4 × 60 × 60 = 0.48 kg. Since
the density of water is 1.0 g cm−3, this mass corresponds to 480 cm3. As a rough guide,
therefore, a 1‐kW fuel cell will produce about 0.5 L of water per hour.
A2.6
Heat Production
Heat is produced when a fuel cell operates. It was noted in Section 2.4, Chapter 2, that
if all the enthalpy of reaction of a hydrogen fuel cell was converted into electrical energy,
then the output voltage would be 1.48 V or 1.25 V if the water product was in liquid
form or vapour form, respectively. Cleary, it follows that the difference between the
actual cell voltage and either of these two voltages represents the energy that is not
converted into electricity, that is, the energy that is transformed into heat.
Since, in most cases, the water is not produced in liquid form, the following analysis
assumes that product water is in the vapour phase and that no account is taken of the
cooling effect of water evaporation. It also means that energy is leaving the fuel cell in
three forms, namely, as electricity, as ordinary ‘sensible’ heat and as the heat of vapourization (latent heat) of water. For a stack of n cells at a current I, the heat generated, in
watts, is given by:
Heating rate nI 1.25 Vc
(A2.20)
In terms of electrical power, in watts, this becomes:
Heating rate
Pe
1.25
1
Vc
(A2.21)
409
411
Appendix 3
Calculation of Power Required by Air Compressor
and Power Recoverable by Turbine in Fuel‐Cell Exhaust
A3.1
Power Required by Air Compressor
The following worked example for compressor power is for a 100‐kW fuel‐cell stack
pressurized at 300 kPa (3 bar). Air is fed to the stack using the Lysholm compressor
whose chart is shown in Figure 12.4, Chapter 12. The air inlet to the compressor is at
100 kPa (1 bar) and 20°C. The fuel cell is operated at an air stoichiometry of 2.0, and the
average cell voltage is 0.65 V, which corresponds to an efficiency of 52% (LHV). The
chart of Figure 12.4, Chapter 12, is used to determine the values of the following
parameters:
●
●
●
●
Required rotational speed of the air compressor.
Efficiency of the compressor.
Temperature of the air as it leaves the compressor.
Power of the electric motor required to drive the compressor.
First, it is necessary to find the mass flow rate of air that will be consumed by the cell
using equation (A2.9):
Air usage
3.58 10
7
2 100 000
= 0.11 kg s
0.65
1
(A3.1)
This value is then converted to the mass flow factor:
Mass flow factor
0.11 293
1
= 1.9 kg s 1 K 2 bar
1.0
1
(A3.2)
Note that the pressure units are in bar, i.e., the same as in the chart given in Figure 12.4,
Chapter 12. The chart can now be used to determine the speed and efficiency of the
compressor, that is, the intercept of a horizontal line drawn from pressure ratio =3 and
a vertical line starting from the x‐axis at a mass flow factor =1.9. The result will be very
close to the 600 rotor speed factor line and the 0.7 ‘efficiency contour’. Thus, the rotor
speed can be taken to be:
600
293 10 300 rpm
Fuel Cell Systems Explained, Third Edition. Andrew L. Dicks and David A. J. Rand.
© 2018 John Wiley & Sons Ltd. Published 2018 by John Wiley & Sons Ltd.
(A3.3)
412
Appendix 3: Calculation of Power Required by Air Compressor and Power Recoverable by Turbine in Fuel‐Cell Exhaust
The efficiency of the compressor and the mass flow rate are used to find the temperature
rise and the compressor power. The former is obtained from equation (12.6), Chapter 12,
namely:
T
293
0.7
30.286 1
155 K
(A3.4)
Since the entry temperature is 20°C, the exit temperature is therefore 175°C. Note that
if the system is a PEMFC, cooling will be necessary. Alternatively, if it is a PAFC, then
the compressor would enable the fuel gas to be preheated.
The power required for the compressor can be determined from equation (12.10),
Chapter 12:
Power 1004
293
0.7
30.286 1
0.11 17.1 kW
(A3.5)
This is the power for the compressor without considering any mechanical losses in
the bearings and driveshafts. The electric motor also will not be 100% efficient — a
reasonable estimate of its power would be about 20 kW. It is important to note the
following:
●
●
The 20 kW of electrical power will have to be provided by the 100‐kW fuel cell, i.e., by
consuming 20% of its output. This parasitic load is a major problem when running
systems at pressure; its importance for PEMFCs is discussed in Section 4.7.2,
Chapter 4.
In this worked example, the assumption is that the air is not humidified, i.e., it has a
low water content. As pointed out in Section 4.4, Chapter 4, the inlet of a PEMFC is
sometimes humidified. This action alters both the specific heat capacity and the ratio
of the heat capacities, γ, and will influence the performance of the compressor.
Humidification, if required, is usually undertaken after compression because the air
is hotter at this stage.
A3.2 Power Recoverable from Fuel‐Cell Exhaust
with a Turbine
The power available from the exit gases of the 100‐kW fuel cell, and recoverable by
using a turbine, can be found as follows.
The mass of the cathode exit gas is increased by the presence of water in the cells, but
since this is the result of replacing O2 with 2H2O, the mass change will be insignificant,
as the mass of hydrogen is so small. The mass flow rate, ṁ , will therefore still be taken
as 0.11 kg s−1. The exit temperature can be estimated as 90°C for a typical PEMFC, and
the entry pressure is 300 kPa (3 bar). The exit pressure must be a little less than this, and
assuming it to be 280 kPa, the mass flow factor can be calculated as:
Mass flow factor
0.11 363
2.8
0.75 kg s
1
K1/2 bar
1
(A3.6)
Appendix 3: Calculation of Power Required by Air Compressor and Power Recoverable by Turbine in Fuel‐Cell Exhaust
The speed and efficiency of the turbine can be determined from the performance
chart given in Figure 12.8, Chapter 12. The intercept on the chart between 0.75 on the
x‐axis and 2.8 on the pressure ratio axis is close to the rotor speed factor line of 5000,
and in the efficiency region of 0.7 or 70%. Consequently, the required rotor speed is
predicted to be:
5 000
363
95 000 rpm
(A3.7)
This very high speed is suitable for directly driving a centrifugal compressor on the
same shaft, but not for a screw compressor. The power available from the turbine can
be obtained from equation (12.10), Chapter 12, i.e.,
1
Power C P
T1
C
P2
P1
1 m
(12.10)
The exit gas is not normal air; it has less oxygen and a changed specific heat capacity.
For engines, standard values are 1150 J kg−1 K−1 for CP and 1.33 for γ. In the case of a fuel
cell, the change in gas composition is not so great, and a value of 1100 J kg−1 K−1 will be
used for CP and 1.33 for γ. The constant (γ − 1/γ) thus becomes 0.275. The temperature
T1 is 363 K, so equation (12.10) becomes:
Power available 100 0.7 363
10.275
1
2.8
0.11
7.6 kW
(A3.8)
The minus sign indicates that power is given out by the turbine. This power is a useful
addition to the 100 kW of electrical output of the fuel cell, but note that it provides less
than half of the power required to drive the compressor, as calculated above.
Furthermore, this example is the best possible result — turbine efficiencies will usually
be somewhat lower than the 0.7 assumed here. As can be seen from the turbine
performance chart in Figure 12.9, Chapter 12, much of the operating region is at
greatly lower efficiency.
413
415
Glossary of Terms
AB5 A range of metal alloys (e.g., LaNi5) capable of undergoing a reversible hydrogen
absorption–desorption reaction.
Absorption The process by which a liquid or gas is drawn into the permeable pores
of a solid material. See: Adsorption; Chemisorption; Physisorption.
Activation energy The energy needed to initiate a chemical reaction. Also known as
‘Arrhenius energy’.
Activation overpotential The overpotential that results from the restrictions
imposed by the kinetics of charge transfer at the electrode|electrolyte interface.
Activity A measure of the ‘effective concentration’ of a species in a reacting system.
By convention, it is a dimensionless quantity. The activity of pure substances in
condensed phases (liquids or solids) is taken as unity. Activity depends principally
on the temperature, pressure and composition of the system. In reactions
involving real gases and mixtures, the effective partial pressure of a constituent gas
is usually referred to as ‘fugacity’. See: Fugacity.
Adiabatic process A process (e.g., expansion of a gas) that takes place without heat
entering or leaving the system; in reversible adiabatic expansion, as a gas cools, its
internal energy is reduced by the amount of work done by the gas on the environment.
Adsorbate A material that has been or is capable of being adsorbed.
Adsorbent A material having capacity or tendency to adsorb another substance.
Adsorption The adhesion of molecules of gases, dissolved substances or liquids
to the surface of solids or liquids with which they are in contact; distinguished
from absorption, a process in which one substance actually penetrates into
the inner structure of the other. Thus, adsorb and adsorbent. See: Chemisorption;
Physisorption.
Alanates Aluminium hydrides of alkali metals or alkaline earth metals, e.g., LiAlH4,
NaAlH4, Mg(AlH4)2.
Alternating current Electric current that flows for an interval of time (half‐period)
in one direction and then for the same time in the opposite direction; the normal
waveform is sinusoidal. Alternating current is easier to transmit over long
distances than direct current, and it is the form of electricity used in most homes
and business. See: Direct current.
Amines Organic compounds that contain nitrogen as a key atom. The compounds are
similar in structure to ammonia, with one or more of the hydrogen atoms replaced
by organic groups, and exhibit a wide range of properties. Amine scrubbing is
commercially used for the removal of carbon dioxide from natural gas.
Fuel Cell Systems Explained, Third Edition. Andrew L. Dicks and David A. J. Rand.
© 2018 John Wiley & Sons Ltd. Published 2018 by John Wiley & Sons Ltd.
416
Glossary of Terms
Amorphous material A solid material in which there is no long‐range order of the
positions of the atoms.
Anaerobic Any process (usually chemical or biological) that takes place without the
presence of air or oxygen.
Anode An electrode at which an oxidation process, i.e., loss of electrons, is occurring.
In a fuel cell, the anode is the negative electrode where hydrogen is consumed.
During electrolysis, the anode is the positive electrode where oxygen is evolved. In a
secondary battery (or cell), the anode is the positive electrode on charge and the
negative electrode on discharge.
Anthracite The highest rank of coal in terms of hydrocarbon content (carbon content
85–95 wt.%) with a glossy, black appearance. It is used primarily for residential and
commercial space heating. Also known as ‘hard coal’. See: Bituminous coal;
Lignite; Peat; Sub‐bituminous coal.
Anthropogenic emission Emission caused, directly or indirectly, by human activities.
The emission of sulfur dioxide due to the use of fossil fuels is an example of a direct
cause of emissions, and the emission of nitrogen oxides from farmland as a function
of fertilizer application is an example of an indirect cause.
Area‐specific resistance The electrical resistance of a sample multiplied by its
geometric area.
Austenitic steel Steel alloys based on austenitic iron (γ‐phase iron). Austenitic
stainless steel contains a maximum of 0.15 wt.% carbon, a minimum of 16 wt.%
chromium and sufficient nickel and/or manganese to retain its elasticity even at
cryogenic temperatures.
Autothermal reforming An energy‐efficient reforming process that uses heat
generated from combining partial oxidation and catalytic steam reforming of a
hydrocarbon feedstock (methane or liquid fuel) in a single step; this can significantly
reduce emissions of carbon dioxide. See: Partial oxidation; Steam reforming.
Balance-of-plant The sum of those components additional to and integrated with the
primary power module of a fuel cell (which is individual fuel cells connected in a stack)
to make up the entire operational system. These components can include a fuel
processor or fuel reformer, power‐conditioning equipment (such as inverters and
voltage controls), motors, compressors, blowers and fans, valves and piping, fuel
storage medium and even conventional batteries complementary to the fuel‐cell stack.
Barrel A measure of crude oil (petroleum), approximately 159 dm3.
Base load The typical minimum electrical power demand placed on a power‐
generating system.
Battery A multiple of electrochemical cells of the same chemistry, connected in series
or in parallel and housed in a single container. (Note that the term is often used to
indicate a single cell, particularly in the case of primary systems.)
BET‐specific surface area The total surface area of a specimen per unit of mass,
usually expressed in m2 g−1, obtained by applying the Brunauer–Emmett–Teller
model to gas adsorption isotherms. Pore volume and pore size-distribution can also
be obtained by this method.
Binder A substance added to the active material of an electrode to enhance
mechanical strength.
Bio‐electrochemical fuel cell A fuel cell that exploits biological species as catalysts
to facilitate the generation of electricity. There are two major classes: ‘enzymatic’
fuel cells and ‘microbial’ fuel cells. The former employ enzymes as catalysts, whereas
Glossary of Terms
the latter employ microorganisms to convert the chemical energy of biofuels
(e.g., glucose, other sugars, alcohols) into electrical energy or hydrogen. Also known
as a ‘biological fuel cell’.
Biofuel A gaseous, liquid or solid fuel that is derived from a biological source.
Biofuel may be in its natural form (e.g., wood, peat) or a commercially produced form
(e.g., ethanol from sugarcane residue, diesel fuel from waste vegetable oils).
Biogas A gaseous fuel of medium energy content, composed of methane (typically
50–60 vol.%) and carbon dioxide, that results from the anaerobic decomposition of
waste matter. Also known as ‘anaerobic digester gas’. See: Anaerobic.
Biomass A collective term used to describe all biologically produced matter at the
end of its life that can be converted to a solid fuel, a renewable liquid fuel (‘biofuel’)
or a gaseous fuel (‘biogas’, such as methane or hydrogen). Biomass can be derived
from forest and mill residues, agricultural crops and wastes (e.g., corn stover, alfalfa
stems, obsolete seed corn, hulls and nut shells, fibre from sugarcane, straw from rice
and wheat), wood and wood wastes (sawdust, timber slash mill scrap), animal
wastes, livestock operation residues, aquatic plants, fast‐growing trees and plants
and municipal and industrial wastes. See: Biofuel; Biogas.
Biophotolysis The storage and use of electrons produced by the first stages of
photosynthesis that can then be used to produce free hydrogen.
Bipolar plate A dense electronic (but not ionic) conductor that electrically connects
the positive electrode in one cell to the negative in the adjacent. The cells are series
connected and so allow the voltage to be built up. Bipolar plates also serve as a
means to distribute fuel or air to the electrodes, to remove reaction products and to
transfer heat. Depending on the type of electrochemical cell, the plate may be made
out of carbon, metal or a conductive polymer (which may be a carbon‐filled
composite). See: End-plate; Flow-field.
Bituminous coal A dense coal (carbon content 45–85 wt.%) that is black, but
sometimes dark brown, often with well‐defined bands of bright and dull material.
It serves primarily as fuel in electricity generation, with substantial quantities also
used for heat and power applications in manufacturing and to make coke or coking
coal, an essential ingredient in making steel. See: Anthracite; Lignite; Peat;
Sub‐bituminous coal.
Butler–Volmer equation The relationship between the current flowing through an
electrode and the potential across the electrode|electrolyte solution interface. At
low overpotentials, it can be very well approximated by a linear relationship, and at
high overpotentials by the Tafel equation. See: Tafel equation.
Calcination Thermal treatment process for solids to induce thermal decomposition
or phase transitions or eliminate volatile components.
Capacitance or Capacity The electric charge stored in a capacitor, measured in
Farads.
Capacitor A device for the temporary storage of electrical charge.
Carbon black An amorphous form of carbon, produced commercially by thermal or
oxidative decomposition of hydrocarbons. It has a high surface area‐to‐volume
ratio, although this ratio is low compared with activated carbon. Often used to
support electrocatalysts.
Carnot cycle The most efficient (‘ideal’) cycle of operation for a reversible heat
engine. It consists of four successive reversible operations, as in the four‐stroke
internal combustion engine, namely, isothermal expansion and heat transfer to the
417
418
Glossary of Terms
system from a high‐temperature reservoir, adiabatic expansion, isothermal
compression and heat transfer from the system to a low‐temperature reservoir,
and adiabatic compression that restores the system to its original state.
See: Adiabatic process.
Carnot efficiency The maximum efficiency with which thermodynamic work can be
produced from thermal energy flowing across a temperature gradient.
Catalyst A substance that increases the rate of a chemical reaction but that is not
itself permanently changed.
Cathode An electrode at which a reduction process, i.e., gain of electrons, is
occurring. In a fuel cell, the cathode is the positive electrode where oxygen is
consumed. During electrolysis, the cathode is the negative electrode where hydrogen
is evolved. In a secondary battery, the cathode is the negative electrode on charge
and the positive electrode on discharge.
Cell voltage The algebraic difference in voltage between the positive and negative
electrodes of an electrochemical cell. Cell voltage usually refers to non‐equilibrium
conditions, that is, when current is flowing through the cell. The term ‘voltage’ is
usually reserved for the case when an electrochemical cell is under consideration,
while the term ‘potential’ is usually reserved for the case when an electrode is
considered. Unfortunately, the two terms are sometimes used interchangeably.
Cermet A composite material composed of ceramic (cer) and metallic (met)
components. A cermet is ideally designed to have the optimum properties of both a
ceramic, such as high‐temperature resistance and hardness, and those of a metal,
such as the ability to undergo plastic deformation. It is typically used as the negative
electrode (anode) in a solid oxide fuel cell.
Chalcogenide A chemical compound that consists of at least one chalcogen ion and
at least one more electropositive element. Although all group 16 elements of the
periodic table are defined as chalcogens, the term is more commonly reserved for
sulfides, selenides and tellurides rather than oxides.
Charge‐transfer coefficient An important parameter in the Butler–Volmer equation
for kinetic treatment of electrochemical reactions. The parameter signifies the
fraction of the interfacial potential at an electrode|electrolyte solution interface
that helps to lower the free energy barrier for the electrochemical reaction.
See: Butler–Volmer equation; Gibbs free energy.
Chelate An inorganic complex in which a ligand is coordinated to a metal ion at two
or more sites.
Chemical potential A form of potential that can be absorbed or released during a
chemical reaction or phase change. For a given component in a mixture, the change
in Gibbs free energy with respect to change in amount of the component, with
temperature, pressure and amounts of other components being constant. See:
Gibbs free energy.
Chemical vapour deposition A chemical process used to produce high‐purity solid
materials, usually as thin deposits. Abbreviated as ‘CVD’.
Chemisorption A process whereby atoms or molecules of the adsorbed substance
(gas or liquid) are held to the surface of a solid material by covalent bonds.
See: Adsorption; Physisorption.
Clathrates A substance in which the molecules of one compound are encapsulated in
lattices or cage‐like structures within another compound. For example, crystalline
Glossary of Terms
clathrates are formed between certain gases (e.g., carbon dioxide, hydrogen sulfide,
methane) and water at low temperatures and high pressures.
Climate change A statistically significant change of climate that is attributed directly
or indirectly to human activity, which alters the composition of the global
atmosphere and is in addition to natural climate variability observed over
comparable time periods. Note that climate is usually defined as the ‘average
weather’, which, in turn, means using statistics to describe weather (temperature,
precipitation and wind) in terms of the mean and variability over a period of time.
The World Meteorological Organization uses periods of 30 years, but periods can
be as short as months or as long as tens of thousands of years. See: Greenhouse
effect; Greenhouse gases.
Coal gas A fuel gas, usually rich in methane, that is produced when coal is heated in
the absence of air (so‐called destructive distillation) or pyrolysis. It is a by‐product
in the preparation of coke and coal tar. Coal gas was a major source of energy in the
late 19th and early 20th centuries and was also known as ‘town gas’. The use of this
gas declined with the increasing availability of natural gas. See: Pyrolysis.
Cogeneration See: Combined heat and power system.
Combined cycle A technology to improve the thermal efficiency of a power station
that uses natural gas as fuel. The gas is first burnt in a gas turbine, which drives a
generator to produce electric power. The waste heat contained in the exhaust gases
is then recovered and used to raise high‐pressure steam, which is expanded through
a steam turbine to drive another electric generator to produce additional power.
Combined‐cycle systems generate electricity in a more efficient and
environmentally sound manner. See: Thermal efficiency.
Combined heat and power system An installation where there is simultaneous
generation of power (either electrical or mechanical) and useful heat (e.g., process
steam) in a single process. Also known as ‘cogeneration’.
Composite The combining or compositing of best performance benefits from two or
more different materials in one component. In a polymer electrolyte fuel cell, for
example, a polymer composite of carbon fibre–epoxy may be used in the bipolar plates;
in a solid oxide fuel cell, the plates and membrane are ceramic, while the interconnects
may be metallic (making the ultimate fuel‐cell stack of composite construction).
Composite membrane An ionically conducting membrane, usually constructed in
the form of a film made from two or more materials, for use in some types of
battery and fuel cell.
Concentration overpotential The potential difference caused by differences in the
concentration of the charge carriers between the bulk solution and the electrode
surface. It occurs when the electrochemical reaction is sufficiently rapid to lower
the surface concentration of the charge carriers below that of bulk solution. The rate
of reaction is then dependent on the ability (‘mass transfer’) of the charge carriers to
reach the electrode surface. Also known as ‘mass‐transport overpotential’ or, less
commonly, as ‘diffusion overpotential’.
Counter electrode An electrode in an electrochemical system that is used only to make
an electrical connection to the electrolyte solution so that a current can be applied
to the working electrode. The processes occurring on the counter electrode are
unimportant; it is usually made of inert materials (noble metals or carbon/graphite)
to avoid its dissolution. Also called an ‘auxiliary electrode’. See: Working electrode.
419
420
Glossary of Terms
Crude oil A mixture of hydrocarbons that exists in the liquid phase in natural
underground reservoirs and remains liquid at atmospheric pressure after passing
through surface‐separating facilities. It occurs in many varieties, distinguished by
specific gravity, concentrations of the component hydrocarbons, volatility, heating
value and sulfur content. Fuels such as motor petrol (also called gasoline), diesel fuel
and jet fuel are derived from crude oil, as well as a variety of materials known as
petrochemicals.
Cryogenic A term applied to low‐temperature substances and apparatus; usually
referring to the temperature range below 77 K.
Current density In an electrochemical cell, the current flowing per unit electrode area.
Cyclic voltammetry See: Voltammetry.
Dead‐end fuel cell A unit fuel cell, or fuel‐cell stack, without fuel and/or oxidant
outlet ports. In a dead‐end operation, all of the reactants fed to the cell, or stack,
are consumed. Care must be taken, however, to allow continuous removal of the
reaction product from the cell/stack. Performance losses are often observed when
one or both reactants are supplied in dead‐end mode. This is caused by non‐
optimum flow distribution, as well as by the accumulation of contaminants or
inert gases.
Density functional theory A quantum mechanical theory used in physics and
chemistry to investigate the electronic structure of many‐body systems, in particular
atoms, molecules and condensed phases.
Desorption Opposite of adsorption, where molecules separate from the solid surface.
See: Adsorption.
Dielectric A substance (solid, liquid or gas) that is a non‐conductor of electricity (i.e.,
an insulator). An electric field in a dielectric substance gives rise to no net flow of
electricity. Rather, an applied field causes electrons within the substance to be
displaced and, thereby, creates an electric charge on the surface of the substance.
This phenomenon is used in capacitors to store charge. See: Capacitor.
Diesel fuel A combustible distillate of petroleum used as a fuel for diesel
(compression ignition) engines; usually the fraction of crude oil that is distilled after
kerosene. See: Crude oil.
Diffusion Net spontaneous and random movement of molecules, particles or ions in
a fluid (gas or liquid) from a region in which they are at a high concentration to a
region of lower concentration, until a uniform concentration is achieved
throughout. The difference in concentration between two such regions is called the
‘concentration gradient’.
Diffusion coefficient The coefficient of proportionality between the flux of a
substance and its concentration gradient.
Diode A solid‐state electrical device that only allows current to flow in one direction.
Direct current Electric current that flows in one direction only, although it may have
appreciable pulsations in its magnitude. It is the form of electricity produced by
electrochemical cells. See: Alternating current.
Direct internal reforming Production of a desired product (e.g., hydrogen) within a
unit fuel cell, or fuel‐cell stack, from a hydrocarbon‐based fuel (e.g., diesel,
methanol, natural gas) fed to the cell, or stack. See: External reforming; Indirect
internal reforming.
Disproportionation A chemical reaction in which a single substance acts as both an
oxidizing and a reducing agent to produce dissimilar substances. For example, carbon
Glossary of Terms
monoxide can decompose over a catalyst to form both solid carbon and carbon
dioxide; this particular disproportionation is known as the Boudouard reaction.
Dissociation, dissociation constant In chemistry and biochemistry, a general
process in which ionic compounds (complexes, molecules or salts) separate or split
into smaller molecules, ions or radicals, usually in a reversible manner. The
equilibrium constant of a reversible dissociation is called the ‘dissociation constant’.
It is the ratio of the product of the concentrations of the dissociated species to the
concentration of the undissociated compound.
Distributed energy A network of power generation, storage and metering/control
systems that allow power to be used and managed in a distributed and small‐scale
manner, thereby placing the supply close to the load, rather than from a large,
centralized power plant, in order to minimize electricity transmission and maximize
waste heat utilization. Also as ‘distributed power generation’ or ‘embedded
generation’.
Drive-train The elements of the propulsion system (including engine, transmission,
driveshaft and differential) that deliver mechanical energy from the power source to
drive the wheels of a given vehicle.
Dye‐sensitized solar cell A photoelectrochemical cell that uses a dye‐impregnated
layer of titanium dioxide to generate a voltage by means of light energy rather than the
semiconducting materials used in most photovoltaic cells. See: Photovoltaic cell.
Electrical double-layer A model of the ionic environment (charge accumulation) at
the interface between an electrode and an electrolyte solution close to it. In general
terms, the structure is comprised of a compact charged layer adjacent to the
electrode surface and a diffuse region of charge extending into the electrolyte
solution. Note that there are several theoretical treatments of the solid|liquid
interface. Also known simply as a ‘double layer’.
Electric vehicle A vehicle that is powered solely by an electrochemical power source,
such as a battery or fuel cell. Power assistance may also be provided by a
supercapacitor.
Electrocatalyst A substance that accelerates the rate of an electrochemical
(electrode) reaction but that is not itself permanently changed.
Electrochemical capacitor A capacitor that stores charge in the form of ions (rather
than electrons), adsorbed on materials of high surface area. The ions undergo redox
reactions during charge and discharge. The device is also known as an
‘electrochemical double‐layer capacitor’, a ‘supercapacitor’ or an ‘ultracapacitor’.
Electrochemical (AC) impedance spectroscopy An investigative technique for the
examination of processes that occur at electrode surfaces. An alternating‐current
(sinusoidal) excitation signal (potential or current), of small amplitude and
covering a wide range of frequencies, is applied to the system under investigation,
and the response (current or voltage or another signal of interest) is recorded.
Given the small amplitude of the excitation signal, data can be obtained without
significantly disturbing the normal operation of the system. By conducting
measurements over a wide range of frequencies, a complex sequence of coupled
processes, such as electron transfer, mass transport and electrochemical reaction,
can often be separated and evaluated. The technique is routinely used to study
electrode kinetics and reaction mechanisms and for the characterization of
battery, fuel cell and corrosion phenomena. The term is abbreviated as ‘EIS’.
See: Impedance.
421
422
Glossary of Terms
Electrode An electronic conductor that acts as a source or sink of electrons that are
involved in electrochemical reactions.
Electrode potential The voltage developed by a single electrode, either the positive
or the negative; usually related to the standard potential of the hydrogen electrode
(which is arbitrarily set at 0 V). The International Union of Pure and Applied
Chemistry defines electrode potential as the voltage of a cell in which the electrode
on the left is a standard hydrogen electrode and the electrode on the right is the
electrode in question. See: Standard hydrogen electrode.
Electrolysis The passage of an electric current through an ionic substance (an electrolyte,
either dissolved in a suitable solvent or molten) that results in chemical reactions at the
electrodes and separation of materials. Note that high‐temperature electrolysis (also
known as steam electrolysis), which is currently being investigated for the production of
hydrogen from water, employs a solid electrolyte of yttria‐stabilized zirconia. See:
Electrolyser.
Electrolyte A chemical compound that ionizes when dissolved or molten to produce
an electrically conductive medium; also solid materials that are conducive due to the
movement of ions through voids, or empty crystallographic positions, in their
crystal lattice structure, e.g., yttria‐stabilized zirconia used primarily in solid oxide
fuel cells. Note that in the case of dissolved materials, it is fundamentally incorrect
to refer to the ‘electrolyte solution’ as the ‘electrolyte’. Nevertheless, the former
terminology has become common practice.
Electrolyser An electrochemical plant designed to effect the process of electrolysis.
Electron An elementary particle with negative electric charge of 1.602 × 10−19
coulombs and a mass of 9.109 × 10−31 kg.
Electron microscope A form of microscope that uses a beam of electrons instead of a
beam of light (as in the optical microscope) to form a large image of a very small object.
Electro‐osmotic drag In certain types of fuel cell, the flux of water that, due to its
attraction to protons, is transported through the electrolyte medium from the
negative electrode (anode) to the positive electrode (cathode). The flux is driven by
the electric field between the electrodes.
Endothermic reaction A chemical reaction that takes place with an uptake of energy
(heat) from the environment. See: Enthalpy; Exothermic reaction.
End-plate A flat metal plate at each end of a fuel‐cell stack, through which tie bolts
are used to compress the cells and cooling plates in the stack, thereby rendering the
whole a continuous electronic conductor. Either the end-plates or the tie bars must
be electrically insulated. If the latter, the end-plate may be the take‐off point for
electric current; otherwise, a separate take‐off on the stack side of each end plate
may be used. See: Bipolar plate.
Energy The ability to do work or produce heat (measured in Joules).
Energy density The accessible stored energy per unit volume of an electrochemical
cell, usually expressed as Wh L−1 or Wh dm−3 (MJ m−3 or kWh m−3 for large storage
facilities). See: Theoretical energy density.
Energy efficiency The ratio of the energy output from a device to the energy input,
usually expressed as a percentage.
Enthalpy A thermodynamic quantity (H) equal to the total energy content of a system
when it is at constant pressure. The gain or loss of energy of a system when it reacts
at constant pressure is expressed by the change in enthalpy, symbolized by ΔH.
Glossary of Terms
When all the energy change appears as heat (Q), the change in enthalpy is equal to
the heat of reaction at constant pressure, i.e., ΔH = Q. The values of ΔH and Q are
negative for exothermic reactions (heat evolved from system) and positive for
endothermic reactions (heat absorbed by system).
Entropy A thermodynamic quantity that represents the amount of energy in a system
that is no longer available to do useful work. When a closed system undergoes a
reversible change, the entropy change (ΔS) equals the energy lost from, or
transferred to, the system by heat (Q) divided by the absolute temperature (T) at
which this occurs, i.e., ΔS = Q/T. At constant pressure, the amount of heat (Q) is
equal to the change in enthalpy (ΔH).
Equivalent weight The weight of a substance that will combine with or displace 1 g of
hydrogen (or 8 g of oxygen) in a chemical reaction. For an element, it is the relative
atomic mass divided by the valency. For a compound, it depends on the reaction
considered.
Eutectic mixture A solid solution that consists of two or more substances and has
the lowest freezing point of any possible mixture of these components. The
minimum freezing point for a set of components is called the ‘eutectic point’. Low‐
melting‐point alloys are usually eutectic mixtures.
Exchange‐current density The current per unit area that flows equally in the
forward and backward directions when an electrode reaction is in equilibrium.
Exergy A thermodynamic property representing the thermal and chemical energy
that is available to do useful work, i.e., it expresses the quality of an energy source.
Specifically, exergy is a measure of the energy difference between some process
state (i.e., pressure, temperature and composition) and a reference state (typically
atmospheric conditions). Exergetic efficiency measures entropy production and
hence represents irreversible losses associated with chemical and thermal
processes.
Exothermic reaction A chemical reaction that takes place with a release of energy
(heat) to the environment. See: Endothermic reaction; Enthalpy.
External reforming The production of hydrogen from a hydrocarbon fuel
(e.g., methanol, natural gas, propane) prior to entry to a unit fuel cell or fuel‐cell
stack. See: Direct internal reforming; Indirect internal reforming.
Fermentation The chemical decomposition of a complex substance, especially a
carbohydrate, into simpler chemical products that is brought about by the action of
enzymes, bacteria, yeasts or molds, generally in the absence of oxygen. May be a
natural process or one promoted or enhanced technically to produce a desired end
product, e.g., corn products, to yield ethanol.
Fischer–Tropsch process A catalysed chemical reaction in which synthesis gas, a
mixture of carbon monoxide and hydrogen, is converted into liquid hydrocarbons
of various forms. Also known as ‘Fischer–Tropsch synthesis’. See: Synthesis gas.
Flow battery A form of rechargeable battery in which the electroactive materials
(usually redox couples) of both electrode polarities are dissolved in a solvent
(usually water) to form electrolyte solutions that are stored externally and pumped
to the cells of the battery during operation. Flow batteries can be rapidly
‘recharged’ by replacing the electrolyte solutions while simultaneously recovering
(‘re-energizing’) the spent material for re‐entry to the battery. Also known as a
‘redox battery’.
423
424
Glossary of Terms
Flow-field The structure of channels in the bipolar plate/separator of some types of
fuel cell that distributes reactants across the surface of a catalysed fuel‐cell
membrane–electrode assembly and also removes the products of the
electrochemical reaction and excess reactants (including inert components of the
reactant inlet stream, e.g., nitrogen). When flow-fields are on one side only, the
plate is called a ‘separator-plate’ or ‘current-collector’ and is not considered bipolar
such as the end-plates in a fuel‐cell stack or the two end-plates in a single cell that is
not integrated into a stack. The most common designs of flow-field are ‘parallel’ (an
array of parallel channels), ‘serpentine’ (gas channels are not straight but contain
bends) and ‘interdigitated’ (a comblike arrangement of discontinuous, so‐called
dead‐ended, channels, with staggered channels respectively connected to the gas
inlet and outlet). The channels are either machined or moulded into flat plates
made of metal, ceramic, graphite or composite. See: Bipolar plate; End-plate;
Membrane–electrode assembly.
Fossil fuels Carbonaceous deposits (solid, liquid or gaseous) that derive from the
decay of vegetable matter over geological time spans.
Fuel‐cell vehicle An electric‐drive vehicle that derives the power for its motor(s)
from a fuel‐cell system.
Fugacity A thermodynamic function that is introduced as an effective substitute for
pressure to allow a real gas system to be considered by the same equations that
apply to an ideal gas. See: Activity.
Gas‐diffusion layer A fuel‐cell component that serves two main functions: it must
permit the passage of gases and be sufficiently conductive to allow the transfer of
electrons. The layer also provides a support for the catalyst, and its structure promotes
the removal of generated water that may hinder (‘flood’) the electrochemical reaction.
The layer is very thin, usually with a thickness of 0.25–0.40 mm and with pore sizes
of 4–50 µm. It is typically made from carbon cloth, carbon paper or Toray paper
(a carbon–carbon composite paper of high strength).
Gasification A special type of pyrolysis where thermal decomposition takes place in
the presence of a small amount of air or oxygen. See: Coal gas; Pyrolysis.
Gasoline Same as Petrol.
Gibbs free energy The energy liberated or absorbed in a reversible process at
constant pressure and constant temperature. Put another way, it is the minimum
thermodynamic work (at constant pressure) needed to drive a chemical reaction
(or, if negative, the maximum work that can be done by the reaction). Thus, the
Gibbs free energy is a thermodynamic quantity that can be used to determine if a
reaction is spontaneous or not. The change in free energy, ΔG, in a chemical reaction
is given by ΔG = ΔH − TΔS, where ΔH is the change in enthalpy and ΔS is the change
in entropy. This is known as the ‘Gibbs equation’. See: Enthalpy; Entropy.
Grain boundary The interface between two regions of a solid that have different
crystal orientations.
Greenhouse effect The trapping of heat by greenhouse gases that allow incoming
solar radiation to pass through the Earth’s atmosphere but prevent the escape to
outer space of a portion of the outgoing infrared radiation from the surface and
lower atmosphere. This process has kept the Earth’s atmosphere about 33°C warmer
than it would be otherwise; it occurs naturally but may also be enhanced by certain
human activities, e.g., the burning of fossil fuels. See: Greenhouse gases.
Glossary of Terms
Greenhouse gases Any of the gaseous constituents of the atmosphere, both natural and
anthropogenic, that absorb and re‐emit radiation at specific wavelengths within the
spectrum of infrared radiation emitted by the Earth’s surface, the atmosphere and
clouds. Greenhouse gases include water vapour, carbon dioxide, methane, nitrous oxide,
halogenated fluorocarbons, ozone, perfluorinated carbons and hydrofluorocarbons.
See: Greenhouse effect.
Half‐cell reaction The electrochemical reaction at an electrode.
Heat-exchanger A device in which heat is transferred from one fluid stream to
another without mixing. Heat‐exchanger operations are most efficient when the
temperature differentials are greater.
Higher heating value of a fuel, HHV The amount of heat released by the complete
combustion of a unit volume or weight of a fuel (initially at 25°C) with all the products
brought back to the original temperature. Thus, the value takes into account recovery
of the latent heat of vapourization of the water formed by combustion and is useful in
calculating heating values for fuels where condensation of the reaction products is
practical (e.g., in a gas‐fired boiler used for space heating). Also known as ‘gross
calorific value’ or ‘gross energy’. See: Lower heating value of a fuel.
Hole The vacancy where an electron would normally exist in a solid. A hole is an
electric charge carrier with a positive charge, equal in magnitude but opposite in
polarity to the charge on the electron. Holes and electrons are the two types of
charge carriers in semiconductor materials; holes are induced into a semiconductor
by adding small quantities of an acceptor dopant to the host crystal. Under the
application of an electric field, holes move in the opposite direction from electrons,
thus producing an electric current. See: Semiconductor.
Hybrid electric vehicle A vehicle that derives part of its propulsion power from an
internal combustion engine and part of its propulsion power from an electric motor
or that uses an internal combustion engine to power a generator to charge a battery
that in turns powers one or more electric‐drive motors. See: Parallel hybrid
electric vehicle; Plug‐in hybrid electric vehicle; Series hybrid electric vehicle.
Hydrogen economy The concept of an energy system based primarily on the use of
hydrogen as an energy carrier and fuel, especially for transportation vehicles and
distributed power generation. See: Distributed energy.
Hydrogen embrittlement A process whereby a metal becomes brittle on exposure to
hydrogen. Embrittlement arises from (i) the formation of metal hydride phases with
different lattice parameters to the non‐hydrided metal, which creates stresses within
the metal lattice, and (ii) recombination of atomic hydrogen to molecular hydrogen
in defects within the metal.
Hydrophilic Having an affinity for water.
Hydrophobic Lacking an affinity for water.
Impedance The analogue of the resistance when applied to alternating current. It is a
measure of the inability of a circuit to carry the electrical current. In many cases, the
impedance varies with the changes in the frequency of the applied electrical
potential due to the properties of the conducting liquid or solid. In electrochemistry,
the impedance of the electrodes is also frequency dependent.
Inconel An austenitic nickel‐based alloy designed for use in high‐temperature
applications. Composed primarily of nickel, chromium, iron and molybdenum.
See: Austenitic steel.
425
426
Glossary of Terms
Indirect internal reforming The reformer unit is separated but adjacent to the
fuel‐cell negative electrode (anode). This arrangement takes advantage of the
close‐coupled thermal benefit from the exothermic reaction of the fuel cell to
support the endothermic reforming reaction. See: External reforming; Direct
internal reforming.
Inductance The magnitude of the capability of a component (inductor), such as a
wire loop or coil, in an electrical circuit to store energy in the form of a magnetic
field. An inductance of 1 H is produced when 1 V is induced by a change in current
of 1 A s−1.
Insulated‐gate bipolar transistor A three‐terminal power semiconductor device,
noted for high efficiency and fast switching. It switches electric power in many
modern appliances, such as fuel cells, electric cars, variable speed refrigerators and
air conditioners. Abbreviated as ‘IGBT’.
Interconnect In a solid oxide fuel cell, the interconnect is either a metallic or ceramic
material that typically lies between each individual cell to allow the cells to be
connected in series and also to allow the passage of fuel and air to the negative
electrode (anode) and positive electrode (cathode), respectively.
Internal resistance The opposition to current flow that results from the various
electronic and ionic resistances within an electrochemical or photoelectrochemical cell.
Internal short-circuit Same as Short-circuit.
Inverter An electronic device that converts low‐voltage direct current to high‐voltage
alternating current.
Ion An atom that has lost or gained one or more orbiting electrons and thus becomes
electrically charged.
Ion‐exchange membrane A plastic film formed from ion‐exchange resin. The utility
of such membranes is based on the fact that they are permeable preferentially only
to either positive ions (cation‐exchange membrane) or negative ions (anion‐
exchange membrane).
Ionic liquid A liquid that essentially contains only ions. In the broad sense, the term
includes all molten salts. Nowadays, however, the term ‘ionic liquid’ is commonly
used for salts whose melting point is below 100°C. In particular, the salts that are
liquid at room temperature are called ‘room‐temperature ionic liquids’ or ‘RTILs’.
Ionization Any process by which an atom, molecule or ion gains or loses electrons.
Ionomer An ionomer is a polymer that comprises repeat units of both electrically
neutral repeating units and a fraction of ionized units. The ionic groups lead to
novel physical properties of the polymer, such as electrical conductivity and
isoviscosity (an increase in ionomer solution viscosity with increase in temperature).
kVAR Unit of reactive power. See VAR.
Latent heat The heat absorbed or released by a substance when it changes state (e.g.,
from solid to liquid, or vice versa) at constant temperature and pressure. The term
Specific latent heat denotes the heat absorbed or released per unit mass of a
substance in the course of its change of state.
Life‐cycle analysis A method for evaluating ‘the whole life of a product’. That is, all
the stages involved, such as raw materials acquisition, manufacturing, distribution
and retail, use and reuse and maintenance, recycling and waste management in
order to create less environmentally harmful products. The process consists of three
parts: inventory analysis (selecting items for evaluation and quantitative analysis),
Glossary of Terms
impact analysis (evaluation of impacts on the ecosystem) and improvement analysis
(evaluation of measures to reduce environmental loads). According to the prime
objective of the exercise, also known as ‘cradle‐to‐grave analysis’, ‘dust‐to‐dust
energy cost’, ‘ecobalance’ and ‘well‐to‐wheel analysis’.
Lignite The lowest rank of coal (carbon content 25–35 wt.%) and used almost
exclusively as fuel for electric power generation; also referred to as ‘brown coal’. See:
Anthracite; Bituminous coal; Peat; Sub‐bituminous coal.
Liquefied petroleum gas Various petroleum gases, principally propane and butane,
stored as a liquid under pressure. Abbreviated as ‘LPG’.
Load The total demand for electrical power on a supply such as a battery, fuel cell or
supercapacitor.
Load following (fuel cell) Method of operating a fuel‐cell system to generate a varying
amount of power, depending upon the load demanded. Due to an inherent lag time
between the change in base load and the response of peripheral balance‐of‐plant
components that support the operation of a fuel cell, load following can be enhanced
with a buffer such as a battery or supercapacitor. See: Balance-of-pant; Base load.
Lower heating value of a fuel, LHV The amount of heat released by the complete
combustion of a unit volume or weight of a fuel assuming that all products remain
in the gaseous state. Thus, the latent heat of vapourization of the water formed by
combustion is not taken into account. The value is useful in comparing fuels where
condensation of the combustion products is impractical or heat at a temperature
below 150°C cannot be put to use. Also known as the ‘net calorific value’. See:
Higher heating value of a fuel.
Luggin capillary A salt bridge with a thin capillary tip at one end that is used to
connect the working and reference electrode compartments of a three‐electrode cell.
Placement of the capillary tip very close to the surface of the working electrode defines
a clear sensing point for the reference electrode and serves to minimize the solution IR
drop. Also known as a ‘Luggin tip’, ‘Luggin probe’ or ‘Luggin‐Haber capillary’.
Mass transport Transfer of materials consumed or formed in an electrode process to
or from the electrode surface. The mechanism of mass transport may include
diffusion, convection and electromigration.
Membrane A layer of material that serves as a selective barrier between two phases
and remains impermeable to specific particles, molecules or substances when
exposed to the action of a driving force. Some components are allowed passage by
the membrane into a permeate stream, whereas others are retained by it and
accumulate in the retentate stream. In a fuel cell, the membrane acts as an electrolyte
(ion exchanger), as well as a barrier film separating the gases in the positive (cathode)
and negative (anode) electrode compartments. See: Ion‐exchange membrane.
Membrane–electrode assembly A core component of the structure of a proton‐
exchange membrane fuel cell that consists of a polymer electrolyte membrane
coated with catalyst–carbon–binder layers (‘electrodes’) and sandwiched by two
microporous conductive layers that function as gas‐diffusion layers and currentcollectors; the assembly is placed between bipolar plates to form the basic unit of a
fuel‐cell stack. Electrochemical reactions occur when a fuel (e.g., hydrogen) and an
oxidant (e.g., oxygen) are applied, respectively, to the negative‐electrode (anode)
and the positive‐electrode (cathode) sides of the assembly. See: Bipolar cell;
Gas‐diffusion layer.
427
428
Glossary of Terms
Microbial fuel cell A bio‐electrochemical system that exploits living microorganisms
as catalysts to facilitate the generation of electricity. Also known as a ‘biological fuel
cell’ or a ‘bio‐electrochemical fuel cell’. Abbreviated as ‘MFC’.
Micro‐electromechanical system The integration of mechanical elements, sensors,
actuators and electronics on a common silicon substrate through microfabrication
technology. Abbreviated as ‘MEMS’.
Micro fuel cells Fuel cells sized for small portable devices such as cell phones,
cameras and laptop computers.
Mixed potential The electrode potential when two electrode reactions occur on the
same electrode surface. The mixed potential has a value in between the equilibrium
potentials of the two electrode reactions; it is a steady‐state phenomenon.
Molar Terminology to denote that an extensive physical property is being expressed
per mole of a substance. (An extensive variable is proportional to the size of the
system, e.g., volume, mass, energy.)
Molarity The concentration of a solution expressed as the number of moles of dissolved
substance dissolved per unit volume of solvent, usually expressed as mol dm−3.
Mole The amount of a substance (in grams) that contains as many elementary units
as there are atoms (6.02 × 1023) in 0.012 kg of the carbon isotope 12C. The elementary
units may be atoms, molecules, ions or electrons.
Mole fraction In a system of mixed constituents, the ratio of the number of moles of
a single constituent in a given volume to the total number of moles of all
constituents in that volume.
Monomer A compound whose simple molecules can be joined together (polymerize)
to form a giant polymer molecule.
Monopolar The conventional method of battery construction in which the
component cells are discrete and are externally connected to each other.
Municipal solid waste Solid domestic/household waste.
Nafion™ A brand name used by DuPont for a series of fluorinated sulfonic acid
copolymers, the first synthetic ionic polymer. The material is resistant to chemical
breakdown, and thus it is useful for membranes in proton‐exchange membrane
fuels cells.
Nernst equation A thermodynamic equation demonstrating that the voltage
developed in an electrochemical cell is determined by the activities of the reacting
species, the reaction temperature and the standard free energy change of the overall
reaction. See: Gibbs free energy.
n‐Type semiconductor A semiconductor in which electrical conduction is due
mainly to the movement of electrons.
Nyquist diagram or Nyquist plot A graphical illustration of the data obtained from
electrochemical impedance spectroscopy. See: Electrochemical (AC) impedance
spectroscopy; Impedance.
Ohmic loss The decrease in the voltage of a fuel cell (or battery) that results from
current flow through the internal resistance.
Oil shale Rocks rich in organic material (kerogen) from which petroleum may be
recovered by dry distillation.
Open‐circuit voltage The voltage of a power source, such as a battery, fuel cell or
photovoltaic cell, when there is no net current flow.
Glossary of Terms
Original equipment manufacturer A confusing term that has two meanings.
Originally, an original equipment manufacturer was a company that supplied
equipment to other companies to resell or incorporate into other products using the
respective resellers’ brand names. A number of companies, both equipment
suppliers and equipment resellers, still use this meaning. More recently, the term is
used to refer to the company that acquires a product or component and reuses or
incorporates it into a new product with its own brand name. Abbreviated as ‘OEM’.
Overpotential The shift in the potential of an electrode from its equilibrium value as
a result of current flow.
Overvoltage The difference between the cell voltage (with a current flowing) and the
open‐circuit voltage. The overvoltage represents the extra energy needed (an energy
loss that appears as heat) to force the cell reaction to proceed at a required rate.
Consequently, the cell voltage of an electrochemical cell (e.g., a rechargeable battery
during discharge) is always less than the open‐circuit value, while the cell voltage of
an electrolytic cell (e.g., a rechargeable battery during charge) is always more than
the open‐circuit value. The overvoltage is the sum of the overpotentials of the two
electrodes of the cell and the ohmic loss of the cell. Unfortunately, the terms
‘overvoltage’ and ‘overpotential’ are sometimes used interchangeably. Moreover,
overvoltage is also referred to as ‘polarization’ of the cell, and overpotential as
polarization of the electrode. This is an ill‐defined and misleading term, as
exemplified by the many different definitions to be found in dictionaries. See:
Open‐circuit voltage.
Parallel connection The connection of like terminals of cells or batteries to form a
system of greater capacity, but with the same voltage.
Parallel hybrid electric vehicle A type of hybrid electric vehicle in which the
alternative power unit is capable of producing motive force and is mechanically
linked to the power-train. See: Power-train; Series hybrid electric vehicle.
Parasitic load Power consumed by the balance‐of‐plant equipment that is necessary
to operate a fuel‐cell system. See: Balance-of-plant; Self‐discharge.
Partial oxidation A combustion process in which just sufficient oxygen is supplied to
oxidize a hydrocarbon fuel to carbon monoxide and hydrogen rather than fully to
carbon dioxide and water. This is accomplished by injecting air with the fuel stream
prior to the reformer. The advantage of partial oxidation over steam reforming of
the fuel is that it is an exothermic reaction rather than an endothermic reaction and
therefore generates its own heat. The hydrogen‐rich gaseous product can then be
put to further use, for example, in a certain types of fuel cell.
Peat A precursor of coal that has industrial importance as a fuel in some regions, for
example, Ireland and Finland.
Permeability The rate of diffusion of gas or liquid through a porous material.
Expressed, for a thin material, as the rate per unit area and, for a thicker material, as
the rate per unit area of unit thickness.
Perovskite Any material with the same type of crystal structure as calcium titanium
oxide (CaTiO3). The general chemical formula for perovskite compounds is ABX3,
where A and B are two cations of very different sizes and X is an anion that bonds to
both the cations. The A atoms are larger than the B atoms. Perovskite compounds
are widely used as positive electrodes (cathodes) in solid oxide fuel cells.
429
430
Glossary of Terms
Petrol Term used in the United Kingdom for a light hydrocarbon liquid fuel obtained
by refining petroleum that is used in most spark‐ignition internal combustion
engines. Other terms for such fuel are ‘gas’, ‘gasoline’ and ‘motor spirit’. See: Crude oil.
Petroleum A collective term for crude oil, natural gas, natural gas liquids and other
related products (both hydrocarbon and non‐hydrocarbon compounds). It is usually
found in deposits beneath the Earth’s surface and thought to have originated from
plant and animal remains of the geologic past. See: Crude oil.
pH A measure of the acidity/alkalinity (basicity) of a solution. The pH scale extends
from 0 to 14 (in aqueous solutions at room temperature). A pH value of 7 indicates a
neutral solution. A pH value of less than 7 indicates an acidic solution; the acidity
increases with decreasing pH value. A pH value of more than 7 indicates an alkaline
solution; the basicity or alkalinity increases with increasing pH value.
Photobiological hydrogen production The production of hydrogen by three classes
of organisms, namely, photosynthetic bacteria, cyanobacteria and green algae.
These organisms use their photosynthetic properties to absorb sunlight and convert
it into chemical energy.
Photoelectrochemical cell Solar cells that extract electrical energy from light, including
visible light. Each cell consists of a photosensitive electrode and a conducting counter
electrode immersed in an electrolyte solution. Some photoelectrochemical cells simply
produce a direct current, while others liberate hydrogen in a process similar to the
conventional electrolysis of water.
Photolysis A chemical reaction (often a decomposition) caused by exposure to light.
Photovoltaic Relating to or designating devices that absorb solar radiation and
transform it directly into electricity.
Photovoltaic cell A semiconductor device for converting light energy into low‐
voltage direct‐current electricity.
Physisorption Adsorption of gases on solid surfaces whereby the bonding is by
means of a weak intermolecular (van der Waals) attraction rather than by chemical
bonding. See: Adsorption; Chemisorption.
Platinum‐group metals The three members of the second and third transition series
immediately preceding silver and gold, i.e., ruthenium, rhodium, palladium and
osmium, iridium, platinum.
Plug‐in hybrid electric vehicle A hybrid electric vehicle with batteries that can be
recharged by connecting a plug to an electric power source. Thereby, it shares the
characteristics of both traditional hybrid electric vehicles, by having an electric
motor and an internal combustion engine, and of battery electric vehicles. See:
Electric vehicle; Hybrid electric vehicle.
Polarization An ill‐defined and misleading term used for overpotential and for
overvoltage. See: Overvoltage.
Porosity The ratio of the accessible volume of a porous body to the total volume,
usually expressed as a percentage. Porosity features such as overall open porosity,
pore shape, size and size distribution are key properties of battery and fuel‐cell
electrodes that significantly influence cell performance.
Potentiostat Electronic hardware required to control a three‐electrode cell and run
most electroanalytical experiments. The system functions by maintaining the
potential of the working electrode at a constant level with respect to the reference
electrode by adjusting the current at an auxiliary electrode. See: Reference
electrode; Working electrode.
Glossary of Terms
Power conditioner The subsystem that converts the direct‐current power from a
fuel‐cell stack subsystem to direct‐current or alternating‐current power that is
required by the application.
Power density The power output of an electrochemical cell per unit volume, usually
expressed as W L−1 or W dm−3.
Power factor The ratio between the total real power (measured in watts or kilowatts)
and the total apparent power (the product of the root‐mean‐square voltage and the
root‐mean‐square current, measured in volt‐amperes or kilovolt‐amperes),
expressed as either a decimal fraction or a percentage.
Power-train The elements of a vehicle propulsion system that include all drivetrain
components plus an electrical power inverter and/or controller but not the battery
or fuel‐cell system. See: Drive-train.
Preferential oxidation A reaction that preferentially oxidizes a gas on a catalyst. For
example, the oxidation of carbon monoxide to carbon dioxide using a heterogeneous
catalyst placed on a ceramic support; a reaction of considerable interest in fuel‐cell
design. Also known as ‘selective oxidation’. Abbreviated as ‘PROX’.
Pressure swing adsorption A technology used to separate some gas species from a
mixture under pressure according to the molecular characteristics of a given gas and
its affinity for an adsorbent material. The process operates at near‐ambient
temperature.
Primary battery (or cell) A battery (or cell) that contains a fixed amount of stored
energy when manufactured and that cannot be recharged after that energy is
withdrawn.
Producer Gas A mixture of carbon monoxide and nitrogen made by passing air over
very hot carbon. The gas is used as a fuel in some industrial processes.
Proton An elementary particle that is stable, carries a positive charge equal in
magnitude to the negative charge of the electron and has a mass of 1.672 × 10−27 kg
(i.e., ~1836 times that of the electron). Also, the nucleus of an ordinary or light
hydrogen atom. Protons are constituents of all atomic nuclei.
Proton‐exchange membrane The polymer‐based component in a proton‐exchange
membrane fuel cell that acts as an electrolyte through which protons, but not
electrons, can pass (to move along the electrode and generate a current), as well as a
barrier film to separate the hydrogen‐rich feed in the positive‐electrode (cathode)
compartment of the cell from the oxygen‐rich negative electrode (anode) side.
Proton‐exchange membranes are also employed in certain designs of electrolysis cells.
Also known as a ‘polymer electrolyte membrane’ or a ‘solid polymer membrane’.
Pyrolysis Thermal decomposition of a substance at elevated temperatures in the
absence of air or oxygen.
Quantum yield For photocells, the fractional number of electrons generated per
photon incident on the cell or the ratio of the number of photon‐induced reactions
occurring to the total number of incident photons.
Rechargeable battery See: Secondary battery (or cell).
Redox battery A battery in which the chemical energy is stored as dissolved redox
reagents. The electrodes are contained in compartments that are typically separated
by an ion‐exchange membrane. See: Flow battery.
Reference electrode An electrode with a reproducible, well‐established potential,
against which potentials of other electrodes can be measured.
Reformate The product of a hydrocarbon reforming process. See: Steam reforming.
431
432
Glossary of Terms
Regenerative braking The recovery of some fraction of the energy normally
dissipated as heat during braking of a vehicle and its return to a battery or some
other energy‐storage device. The process of slowing a vehicle involves drawing
kinetic energy into a motor so that it acts as an electric generator and thereby
exerts a rotational drag on the wheels. Most hybrid electric vehicles employ
regenerative braking.
Regenerative fuel cell A type of fuel cell in which the chemical reactants undergo
reversible reactions, such that the cell may be recharged with a separate power
source if desired. For example, the hydrogen–oxygen fuel cell may be recharged for
the production of hydrogen via water electrolysis. Also called a ‘reversible fuel cell’.
Abbreviated as ‘RFC’. See: Unitized regenerative fuel cell.
Relative humidity The ratio of the actual amount of moisture in the air to the
amount needed for saturation at the same temperature.
Renewable energy Forms of energy (such as geothermal heat, hydropower, sunlight,
tidal energy, wave power, wind and organic matter) that flow through the Earth’s
biosphere and are available for human use indefinitely, provided that the physical
basis of their flow is not destroyed. Also known as ‘renewables’.
Reversible potential The potential of an electrode when there is no net current
flowing through the cell.
Reversible voltage The difference in the reversible potentials of the two electrodes
that make up the cell. See: Reversible potential.
Round‐trip efficiency Same as Energy efficiency.
Saturated calomel electrode A reference electrode based on the reaction between
elemental mercury and mercury (I) chloride (Hg2Cl2, ‘calomel’). The aqueous phase
in contact with the mercury and the mercury (I) chloride is a saturated solution of
potassium chloride in water. The electrode is normally linked via a porous frit
(‘salt bridge’) to the solution in which the other electrode is immersed. The
equilibrium electrode potential is a function of the chloride concentration in the
internal electrolyte solution. At 25°C, the potential of the saturated calomel
electrode is +241.2 mV versus the standard hydrogen electrode.
Secondary battery (or cell) A battery (or cell) that is capable of repeated charging
and discharging. Also known as a ‘rechargeable battery (or cell)’.
Selective oxidation See: Preferential oxidation.
Semiconductor A solid‐state crystalline material that has a value of electrical
resistivity intermediate between those of metals and insulators. The conductivity of
semiconductors can be controlled by adding very small amounts of foreign elements
called ‘dopants’. Conductivity is facilitated not only by negatively-charged electrons
but also by positively-charged holes, and it is sensitive to temperature, illumination
and a magnetic field. See: Hole.
Sensible heat The heat absorbed by a substance that gives rise to an increase in
temperature of the substance. See: Latent heat.
Separator An electronically non‐conductive, but ion‐permeable, material that
prevents electrodes of opposite polarity from making contact.
Sequestration The capture of carbon dioxide from streams of mixed gases and its
subsequent indefinite storage. Also known as ‘carbon capture and storage’.
Series hybrid electric vehicle A type of hybrid electric vehicle that runs on battery
power like a pure electric vehicle until the batteries discharge to a set level, when an
Glossary of Terms
alternative power unit turns on to recharge the battery. See: Parallel hybrid
electric vehicle.
Short circuit The direct connection of positive and negative electrodes either
internal or external to the battery.
Short‐circuit current Current flowing freely through an external circuit that has no
load or resistance; the maximum current possible.
Sintering A method for making objects from powder by heating the material (below
its melting point — solid‐state sintering) until its particles adhere to each other.
Sintering is employed in the manufacture of membranes for solid oxide fuel cells.
Sol–gel synthesis A method of preparing single or mixed oxides at low temperature
that involves the formation of a sol (a colloidal suspension or solution of the
precursor), which is converted to a gel (a continuously linked network of oxide
with a continuous interstitial liquid phase) before drying to form a xerogel.
See: Xerogel.
Solid electrolyte A solid‐state ion conductor in which the electrical conductivity is
due to the movement of ions (cations or anions) through voids or interstitial spaces in
the lattice structure. Also known as ‘fast ion conductors’ or ‘superionic conductors’.
Specific energy The accessible stored energy per unit mass of an electrochemical
cell, expressed as MJ kg−1, Wh kg−1 or kWh kg−1. See: Theoretical specific energy.
Specific heat The quantity of heat that unit mass of a substance requires to raise its
temperature by one degree, expressed as J kg−1 K−1.
Specific power The power output of an electrochemical cell per unit weight, usually
expressed as W kg−1.
Specific surface area Total surface area of a material divided by the mass of the
material, usually expressed as m2 g−1. See: BET‐specific surface area.
Sputtering A method of depositing a thin layer of one material on to a substrate.
A target material is bombarded by charged particles (typically argon) that dislodge
atoms from the target and deposit them on a substrate. The technique is a form of
physical vapour deposition.
Standard conditions for temperature and pressure A standard set of conditions for
experimental measurements to allow comparisons to be made between different
sets of data. The International Union of Pure and Applied Chemistry (IUPAC) has
established two standards: (i) standard temperature and pressure, abbreviated as
‘STP’, specifies a temperature of 273.15 K and an absolute pressure of 100 kPa (1 bar)
and (ii) standard ambient temperature and pressure, abbreviated as ‘SATP’, specifies
a temperature of 298.15 K and an absolute pressure of 100 kPa (1 bar). By contrast,
the standard formulated by the National Institute of Science and Technology (NIST)
in the United States stipulates a temperature of 293.15 K and an absolute pressure of
101.325 kPa (1 atm), abbreviated as ‘NTP’.
Standard electrode potential The reversible potential of an electrode with all the
active materials in their standard states. The standard states usually adopted by
electrochemists specify an absolute pressure of 101.325 kPa (1 atm) for gases and
unit activity for elements, solids and 1 mol dm−3 solutions — all at a temperature of
298.15 K. See: Reversible potential.
Standard hydrogen electrode A standard reference electrode, usually consisting of a
platinum electrode coated with platinum black that is bathed with a stream of
hydrogen gas bubbles and immersed in a solution of hydrogen ions (typically
433
434
Glossary of Terms
sulfuric acid). Its potential is declared to be 0 V at all temperatures when the activity
of all species is unity. A zero point is needed since the potential of a single electrode
cannot be measured — only the difference of two electrode potentials is measurable.
All electrode potentials are expressed on this ‘hydrogen scale’. In practice, a unit
concentration (rather than unit activity) of hydrogen ions and unit pressure
(rather than unit fugacity) of hydrogen gas are used. Other reference electrodes
(e.g., calomel or silver|silver chloride) are often employed, but the measured
electrode potentials can be converted to the hydrogen scale. See: Activity; Fugacity;
Saturated calomel electrode.
Steam reforming The reaction of fossil fuels with steam at high temperature to
generate a mixture of hydrogen and carbon monoxide (‘synthesis gas’). It is the
dominant method of commercial hydrogen production and is based on reacting
methane (natural gas) with water; carbon dioxide is formed as a by‐product.
See: Synthesis gas.
Steam‐to‐carbon ratio The number of moles of water per mole of carbon in either
the reformate or the fuel streams. This term is used when steam is injected into the
reformate stream for the water–gas shift reaction or into the fuel for steam
reforming. See: Reformate; Steam reforming; Water–gas shift reaction.
Stoichiometric ratio The perfect oxidant to fuel ratio in a reaction such that all of
the oxidant exactly reacts with all of the fuel.
Stoichiometry The branch of chemistry concerned with the exact or fixed relative
proportions of elements in a chemical compound or of reactants to produce a
compound. For example, in carbon dioxide, the stoichiometric ratio of carbon atoms
to oxygen is 1 : 2. Stoichiometric amounts satisfy a balanced chemical reaction with
no excess of reactants or products.
Sub‐bituminous coal A medium‐soft coal (carbon content 35–45 wt.%) with
properties that range from those of lignite to those of bituminous coal. It is used
primarily as fuel for electricity generation and is an important source of light
aromatic hydrocarbons for the chemical synthesis industry. See: Anthracite;
Bituminous coal; Lignite; Peat.
Substitute natural gas A fuel gas with similar properties to those of natural gas and
that can be produced from fossil fuels (such as lignite coal) or from biofuels. It can
be distributed in the natural gas grid, provided it fulfils the strict criteria for net gas
feeding. Abbreviated as ‘SNG’.
Sustainable energy (‘sustainability’) A set of energy technologies that meets
humanity’s needs on an indefinite basis without producing irreversible
environmental effects. (Note that various definitions exist in the literature, but they
all convey the same message.)
Synthesis gas A gas mixture that contains varying amounts of carbon monoxide and
hydrogen. Examples of production methods include the steam reforming of natural
gas or liquid hydrocarbons, the gasification of coal or biomass and some types of
waste‐to‐energy gasification processes. Also known as ‘syngas’. See: Steam reforming.
Synthetic natural gas Methane produced by the catalytic reaction of carbon
monoxide with hydrogen or from coal by reaction with hydrogen.
Tafel equation The relationship between the current flowing and the overpotential of
an electrode. A plot of electrode potential versus the logarithm of current density is
called the ‘Tafel plot’ and the resulting straight line the ‘Tafel line’. The slope
Glossary of Terms
provides information on the mechanism of the electrochemical reaction, and the
intercept on the current axis (abscissa) provides information on the rate constant
(and exchange‐current density) of the reaction. See: Butler–Volmer equation;
Exchange‐current density.
Tape casting A method for the production of thin, flat ceramics. The ceramic powder
is blended with a liquid and binder (this mixture is referred to as a ‘slip’) and then
deposited on a moving flat surface that is passed under a flat blade to create a
continuous tape. The tape is heat treated to remove the liquid phase and binder and
sintered to promote bonding between the ceramic particles.
Theoretical energy density The energy output of an electrochemical cell referred to
the volume of only the active materials and a 100% utilization of these materials,
expressed as Wh L−1 or Wh dm−3.
Theoretical specific energy The energy output of an electrochemical cell referred to
the weight of only the active materials and a 100% utilization of these materials,
expressed as Wh kg−1.
Thermal efficiency For a heat engine, the ratio of the useful work done by the engine
in a given time interval to the mechanical equivalent of the heat energy supplied in
the steam or fuel during the same time interval.
Thermal expansion coefficient A parameter used to express the dimensional
response of a given solid material to a given unit change of temperature, specifically,
the ratio of change in dimensions to original dimensions per degree rise in
temperature.
Thermochemical (hydrogen) cycle A multistep chemical reaction that sums to the
overall production of hydrogen (and oxygen) by water decomposition; the Carnot
efficiency of the step performed at the highest temperature places a theoretical limit
on the overall hydrogen production efficiency of such a cycle. See: Carnot efficiency.
Three‐phase boundary The gas|electrocatalyst|electrolyte interface formed in a
fuel‐cell electrode such that the electrocatalyst has simultaneous contact with the
reactant gas, the ionic conductor (electrolyte) and the electron conductor. The
electrochemical reactions occur at these points of simultaneous contact. Also
known as a ‘triple‐phase boundary’.
Tortuosity The distance a molecule or ion must travel to get through a substance film
divided by the thickness of the substance.
Town gas See: Coal gas.
Transformer An electrical device that steps up or steps down voltage. Transformers
work with alternating current only.
Transport number The fraction of the total current flowing in an electrolyte phase
that is carried by a particular ion. Also known as ‘transference number’.
Triple‐phase boundary See: Three‐phase boundary.
Unitized regenerative fuel cell A unitized regenerative fuel cell based on a proton‐
exchange membrane that can perform the electrolysis of water in regenerative mode
and function in the other mode as a fuel cell by recombining oxygen and hydrogen
to produce electricity. Abbreviated as ‘URFC’. See: Regenerative fuel cell.
Valence A number that indicates the combining power of one atom with others, that
is, the number of other atoms with which it can combine.
VAR (volt‐amperes reactive) A unit of reactive power in a circuit that is carrying a
sinusoidal current. A VAR equals the amount of reactive power in the circuit when
435
436
Glossary of Terms
the product of the root‐mean‐square value of the voltage (volts) by the root‐mean
value of the current (amperes), and by the sine of the phase angle between the
voltage and the current, equals 1.
Viscosity The resistance of a fluid to shear forces and hence to flow.
Voltage The difference in potential between the two electrodes of a cell or the
two terminals of a battery.
Voltammetry An electrochemical measuring technique used for electrochemical
analysis, for the determination of the kinetics and mechanism of electrode reactions,
and for corrosion studies. ‘Voltammetry’ is a family of techniques with the common
characteristics that the potential of the working electrode is controlled (typically
with a potentiostat) and the current flowing through the electrode is measured.
‘Linear‐sweep voltammetry’ involves scanning the potential linearly in time
(the plots are known as ‘voltammograms’). ‘Cyclic voltammetry’ is a linear‐sweep
voltammetry with the scan continued in the reverse direction at the end of the first
scan; this cycle can be repeated a number of times. In alternating current (AC)
voltammetry, an alternating voltage is superimposed on the direct‐current ramp.
Water-gas A mixture composed primarily of hydrogen and carbon monoxide
produced by passing steam over incandescent carbon, usually from anthracite coal
or coke. The reaction is strongly endothermic but may be combined with the
exothermic reaction for producer gas. Used for lighting (mainly during the 19th to
early 20th century) and as a fuel (well into the 20th century). See: Producer gas;
Steam reforming; Water-gas shift reaction.
Water-gas shift reaction The reaction of water–gas with steam to yield hydrogen
and carbon dioxide.
Well‐to‐wheels analysis See: Life‐cycle analysis
Working electrode The electrode in an electrochemical system on which the
reaction of interest is taking place. The kinetics and mechanism of the reaction may
be under investigation, or the reaction occurring on the working electrode may be
used to perform an electrochemical analysis of the electrolyte solution. The
electrode can serve either as a positive or a negative according to the applied
polarity. See: Counter electrode.
Xerogel A solid formed from a gel by drying with unhindered shrinkage. Xerogels
usually retain high porosity (25%) and high surface area (150–900 m2 g−1), along with
a very small pore size (1–10 nm).
X‐ray photoelectron spectroscopy A quantitative technique used to determine the
elemental composition, empirical formula, chemical state and electronic state of the
elements on the surface of a material. The analysis is performed under ultrahigh
vacuum conditions. Spectra are obtained by irradiating a specimen with a beam of
X‐rays while simultaneously measuring the kinetic energy and number of electrons
that escape from the top 1–10 nm of the material under investigation. Also known
as ‘electron spectroscopy for chemical analysis’ (ESCA). Abbreviated as ‘XPS’.
Zeolite Any one of a family of hydrous aluminium silicate minerals with a cage‐like
molecular structure; used chiefly as molecular filters, ion‐exchange agents and
catalysts (either used directly, e.g., in petroleum refineries, or loaded with catalysts
for other chemical reactions).
437
Index
a
Absorption 74, 137, 155, 204, 276, 293,
313, 334–7, 345
Acetaldehyde 157, 171
Acetonitrile 313
Acid electrolyte 7–8, 50, 58, 161, 163,
172, 174, 307, 341
Acid fuel cell 7, 69, 136, 172
AC impedance 65
spectroscopy 61–2, 64, 190
AC power 371, 378
source 373
Activation losses 46, 60, 66–8, 312
Activation overpotential 48–9, 51–2, 59,
65–7, 153–4, 162–3
Activity 36–7, 86–8, 172–3, 205–6,
254–5, 415, 433–4
electrocatalytic 213
Acumentrics, Inc. 255–6
AC voltage 62, 375, 378, 379
Adsorption 88, 176, 293, 337, 345, 415,
418, 420, 430
dissociative 85, 161, 171
pressure swing 292, 431
Advantica Technologies Ltd. 301
Air‐breathing fuel‐cell stack 106
Air compressor 116, 118, 310, 356, 411–13
Aircraft 169, 268
Air electrode 20, 57, 82, 100, 149, 158,
205, 249
Air‐electrode supported (AES) 249
Air flow 94, 98, 124, 364
Air humidity 96, 100
Air usage 357, 405–7, 411
Air utilization 192
Alanates 337, 342, 415
Algae 268, 272, 319
Alkali metal
carbonates 207, 211
hydrides 325–6, 339–40, 415
Alkaline electrolyser 150, 306, 310,
312, 337
Alkaline electrolyte 8–9, 17, 20, 69, 150,
155, 162, 170, 172, 176, 305
Alkaline fuel cell (AFC) 6–8, 17–18,
69–70, 135–47, 149–56, 160, 178–9,
207, 305–6, 341–2
Apollo 7, 17–18, 136, 140–2, 153–4
Bacon 147
catalysts 149
dissolved fuel 142–3
membranes 144, 146
stacks 140, 149–51
systems 18, 139, 142, 310
Alkaline proton‐exchange membrane
fuel cell (APEMFC) 144
Alkanes 266–7
Alkylbenzene 73
Allied Signal Aerospace Co. 257
Allis‐Chalmers 139
Alloys, AB5, 180, 334, 415
Alstom 395
Altergy 123
Alternating current (AC) circuit 368, 378–9
Fuel Cell Systems Explained, Third Edition. Andrew L. Dicks and David A. J. Rand.
© 2018 John Wiley & Sons Ltd. Published 2018 by John Wiley & Sons Ltd.
438
Index
American Gas Association 7, 205
Amide 341
Amine 145, 340
scrubbing 415
Ammonia 87, 142, 198–9, 264–5, 273–4,
295–6, 325, 343, 389, 415
borane 340–1
Anaerobic 273, 319
Analogue‐to‐digital converter 377
Anion‐exchange membrane (AEM)
fuel cells 144–6, 155
Anionic membranes 175, 182
Anion‐permeable membrane 180–1
Anode 3, 7–9, 11–16, 142–4, 157–67,
179–81, 207–10, 219–22, 233–6
alloy 213
catalyst 88, 115, 162–3, 166, 172–5,
177–8, 180, 198, 204, 267
ceria cermet 245
composite 246
copper cermet 284
cylindrical 249
palladium 180
reaction 160, 162, 167, 170, 172, 174,
203, 213, 282–3
Anode‐supported electrolyte cells 244
Anthracite 269, 416–17, 427, 434
Apparent power/kVA 379–90
Area specific resistance (ASR) 55, 57,
242–3, 416
Arrhenius 415
Asbestos 142
Atomic mass unit (amu) 29
Austenitic steel 248, 425
Auto‐ignition 264, 330
Autothermal 173, 301, 416
reforming 285–6, 303
Auxiliary power units (APUs) 132, 183,
255, 304
Avogadro’s law 53
Avogadro’s number 29
Axial fans 363–4
Axial flow design 352
b
Bacon 6, 135, 137, 147, 153, 155, 397
Bacteria 169, 264, 273, 292, 318, 423
anaerobic 320
photosynthetic 430
Baker 132, 230, 275, 344
Balance‐of‐plant (BoP) 114, 128,
351, 389
components 23, 196, 228, 291, 376
Ballard, Geoffrey 7, 125, 126
Ballard Power Systems (BPS) 16, 70–1,
77, 125, 300, 304, 395, 398
Barium zirconate 294
Barriers 25, 72, 306
bubble 213
selective 427
Battery 1, 4, 6, 31, 182, 350–1, 387, 398,
419, 421, 425
bromine 20, 22
cadmium 349
chargers 122, 157, 380
gaseous voltaic 4
lithium 126–7, 182
metal hydride 126
primary 431
redox 423
secondary 31, 131, 416, 418, 431–2
vanadium 21
BIMEVOX 241
Binder 145, 148, 198, 210,
416, 434–5
mixture 213
organic 212
Bioethanol 169
cellulosic 386
Biofuels 168, 263, 265, 268, 272–4, 275,
318, 323, 417, 434
Biogas 223, 230, 272–4, 417
anaerobic 225
compositions of 274
Biological
digestion 273
fuel cells 23, 417, 428
hydrogen generation 318
shift reaction 320
Biomass 169, 272–4, 321, 386–7, 417,
434
Biophotolysis 319, 417
Bio‐reactor 320
Bio‐separation processes 145
Index
Bipolar plates 11–16, 54–5, 89–90,
102–3, 105–12, 114, 149, 151,
213, 237–8
AFC stacks 152
alkaline electrolyser 306
construction 110
designs 72, 215
flow‐field plate 82
metal 109
multilayer 199–200
ribbed 199
Bloom 257
Blowers 120, 124, 137, 224, 306, 351,
353, 363–4, 416
BMW 305, 324, 333–4
Bode plot 62, 64
Boilers 195, 201, 227, 362
gas‐fired 425
Borohydride 144, 179–81, 342–4
alkali metal 342
decomposition 179
Boudouard reaction 221, 245, 280–1, 421
Bromine 20, 22
Buckminsterfullerene 345–6
Bus 71, 125–6, 130, 264, 267–8, 304, 328,
366, 375, 395
articulated city 139
diesel 127
Butane 268, 272, 324, 427
Butler–Volmer equation 49, 393,
417–18, 434
c
Cadmium telluride 313
Calcia‐stabilized zirconia (CSZ) 239
Calomel 60, 432, 434
Calorific value 27
Caltex Oil Corporation 206
Capacitance 59–60, 65–6, 367, 373,
380, 417
double‐layer 65, 421
Capacitative discharges 290
Carbohydrate 319, 423
Carbon 16–17, 163–4, 231–4, 277–8
activated 276, 282, 417
bipolar plates 107, 110
capture and storage 432
catalyst 163, 198
cloth 90
corrosion 89, 132, 199, 205
deposition 173, 219, 221, 246, 280–1,
301, 390
electrodes 148, 345
fibre 89, 108, 148, 198, 419
filaments 281
functionalizing 90
graphitic 107
graphitized 129
multiwalled 84, 346
nanofibres 344–6
nanomaterial 289
nanotubes 87, 91, 163, 345
paper 90, 198, 320
turbostratic 234
Carbonates 70, 137, 142, 149, 210–11,
213–14
deactivation 150
electrolyte 197
mixtures 212
molten 233
poisoning 219
precipitates 144
Carbon dioxide 161–2, 207–9, 270–1,
273–4, 319–21, 431–2, 434
atmospheric 318
emissions 25, 416
removal 138, 155, 415
separation 23, 230
Carbon monoxide 202–4, 209, 235–6,
273, 403–4, 408–9, 434
adsorbed 171
dehydrogenase 320
fuel cell 209, 403, 404
tolerance 76
Carnot cycle 5, 32, 417
efficiency 418, 435
limit 33, 34
Catalyst 52, 61, 114–15, 143, 147–8,
149–50, 163–7, 168–74, 176, 200,
218–20, 275–8, 280–1, 288–92,
338–41, 380
carbon‐supported 89, 148
chromium 290, 303
cobalt 276
439
440
Index
Catalyst (cont’d )
composite 131
copper 290–1
degradation 88, 129, 132, 219, 301
layers 68, 78, 82–3, 85, 92–4, 164,
198–200, 392–3
molybdenum oxide 276
nickel alloy 205
non‐precious metal 84, 86, 167, 172, 205
overpotential 100, 117, 221–2, 230,
247, 393
palladium 177
particles 10, 83, 90, 285
photoelectrochemical 314
platinum‐based 82, 87, 136, 286
poison 189, 219
precious metal 187, 235, 291, 340
reactions 143–4, 148, 207–9, 236
thorium‐based 273
titanium 342
Catalytic partial oxidation (CPO) 285,
303–5
Catalytic rich gas (CRG) 269
Cathode 7–9, 11–16, 50–2, 159–67,
195–9, 207–10, 233–8
Cation 2–3, 180–1, 241, 245, 429, 433
Cation‐permeable membrane 181
Cella Energy Ltd. 344
Ceramatec Inc. 257
Ceramic Fuel Cells Ltd. (CFCL) 258
Ceramic materials 8, 16, 19, 212, 226,
235–7, 294, 426
Ceramic processing techniques 240
Ceramic substrate 258, 293
extruded 257
Cerate 294
Ceria 172, 240–1, 245–6, 291, 305, 317
cerium gadolinium‐doped (CGO) 241,
244, 245
samarium‐doped 241
Cerium 180, 215
Cermet 235, 244–6, 294, 418
Chalcogenide 87, 167, 418
Channels 13–15, 90, 105–6, 114, 301, 393
cooling water 200
flow‐field 82, 108, 218
staggered 424
Charge 65, 123, 127, 370–2, 380, 382–3,
405–6, 416, 418, 429
accumulation 421
carriers 69, 419, 425
distribution 295
electric 240, 425
electrical 58, 417
flow 8–9
state 22, 182
transfer 48–9, 58, 64, 382, 415
Chelates 86, 418
Chemical looping 314, 317–18
Chemical structure 72–3, 78
Chemical vapour deposition (CVD) 240,
294, 345, 418
Chemical vapour infiltration 108
Chemisorption 415, 418, 430
Chloride concentration 432
Chloride ion 305
Chloromethylation 145
Chlorophyll 319
Choking limit 362
Chromite 247
Chromium 147, 205, 213, 237, 248, 416
photocatalyst 313
poisoning 238
steels 330
Clathrate 347, 418–19
Clean Urban Transport Europe
(CUTE) 298
ClearEdge Power 397–8
Climate 399, 419
Coal 5–6, 34, 225, 234, 263, 265,
268–71, 284, 317, 323, 419, 424,
429, gas 19
anthracite 436
bituminous 269, 416–17, 427, 434
brown 427
coking 417
gasification 6, 233, 273, 434
hard 269, 416
lignite 434
tar 419
Coal‐bed methane (CBM) 270–2
Coal seam gas (CSG), see Coal‐bed
methane (CBM)
Coating 108–9, 168, 215, 294
Index
Cobalt 85–7, 129, 148, 150, 172, 205, 215,
273, 276
Cobalt chloride 342
Cobalt oxide 291, 314
Co‐flow 39, 111, 191, 217, 284
Cogeneration 254, 286, 378, 389,
397, 419
Coke 269, 417, 419, 436
Collector 367, 370
current 11, 112–14, 149, 213, 215–17,
244, 255, 258, 312–13, 424
Combined‐cycle systems 253, 419
Combined heat and power, see
Cogeneration
Combustion 5, 32, 193, 249, 251, 264,
269, 296, 338, 387, 390–1
catalyst 300
chemical looping 317–18
enthalpy 391
partial 286, 302–3
pressurized 299
products 299, 317, 427
reaction 300
Combustor 208, 231, 298
Commercialization 18, 25, 72, 205, 225,
254–5, 384, 394
sustained 182
Commonwealth Scientific and Industrial
Research Organisation (CSIRO)
258, 383
Compact regenerative reformers 299
Composite curves 196
Composite cylinders 159, 328–9,
347–8
Composites 108, 346, 419
cylinder 137, 159
hydrogen storage 330
metal‐ceramic 293
mica‐based 248
Compressed gas 158, 310, 324, 327–8,
337, 348, 356, 385
Compression 107, 116, 288, 310, 326–7,
332, 354–5, 412
isothermal 418
mechanical 132
moulding 108
ratios 353
Compressor 23–4, 98, 118–20, 351–6,
358–9, 361–2, 411–13
anode recycle 124
axial 354–6
centrifugal 353–4, 358–9, 361, 364
efficiency 118, 355
multistage 358
recycle 193
rotating vane 354
screw 353–4, 413
Concentration gradient 293, 420
Concentration losses 46, 94
Concentration overpotential 59, 131, 419
Conduction 214, 239
band 313
electronic 241, 244
Conductivity 75–6, 80–1, 137, 148, 152,
202, 240–3, 245, 254, 432
mixed 247
proton 294
Conductor 235, 290, 417
fast oxygen‐ion 240
mixed ionic–electronic conducting
(MIEC) 243, 246, 284, 294–5, 301
oxide‐ion 238, 243–4
proton 81
Contaminants 124, 137, 270, 302, 420
Control
circuits 370–1
elements 392
switches 373
valves 23, 164, 392
Controller 124–5, 377, 381, 431
programmable logic controller
(PLC) 129
Convection 104, 165, 299, 427
Conversion efficiency 177, 182, 307,
319–20
Cooling 102, 107–9, 114, 116, 197,
290–1, 296–7, 351, 354, 356
air 15, 104–6, 138, 141, 194, 231, 363
channels 105, 107, 128, 200
coils 306
curves 195
fluid 14, 105, 107, 112, 142, 187, 200
Copper 1, 86, 109, 213, 246, 280, 306, 319
cermet 284
441
442
Index
Corrosion 88, 107, 109–10, 114, 144, 201,
215, 230, 237, 267
resistance 142, 210
Cost 25, 39, 70, 72, 85, 98–9, 120–1,
149, 168, 193, 213, 219–20, 291,
330, 394–5
life‐cycle 388
reduction 293
target 158, 385
Coulombs 29–30, 422
Counter‐current 299
Counter‐electrode 60
Counter‐flow 39, 111, 217
Cracks 212, 289, 330
Creepage 210, 219
Cross‐flow configurations 39, 200
Crossover 46, 52, 54, 85, 143, 165–7, 173,
176–8, 182, 197
Crude oil 266–70
Cryogenic 420
liquid 265, 326, 331
storage vessels 153
Crystal structure 241, 247, 293, 429
fluorite 240
Current density 24, 43, 46–7, 49–50, 67,
70–1, 139, 148–50, 201–2, 204, 312
distribution 39, 112, 191
limiting 57
local 112, 188, 190
Current‐interruption 65
Current losses 54
Current path 113, 138, 149, 253, 367,
369, 371
Cyanobacteria 319–20, 430
Cyclic voltammetry (CV) 60–1, 84–5,
355, 420, 436
Cyclohexane 267, 289, 338–9
Cyclopentane 267
d
Daimler 125–7, 157, 344, 395
Dark fermentation 321
Davy, Sir Humphry 1, 2
Dead ended 115–16, 124, 420
Decomposition 5, 280, 323, 334, 345, 430
anaerobic 417
chemical 423
direct 179
oxidative 417
rate 343
Degradation 25, 93, 129, 137, 144, 146,
197, 199, 219, 237, 245, 248, 337,
378, 380
acid‐catalysed 77
mechanical 129, 245, 252, 293
permanent 247, 295
progressive 310
Delphi Automotive LLP 257
Density functional theory (DFT)
394, 420
Desorption 293, 415
cycles 137, 334–7
pressure 335
reaction steps 175
Desulfurization 23, 189, 275, 296–8, 390
Desulfurizer 193, 224, 227, 269, 282,
297, 299
Detergents 73, 267
Devolatilization 269
Dewar 324, 331
Dew point 95, 102
Diaphragm pump 365
Dibenzyl toluene 339
Diborane 325
Dicks, A. L. 63–4, 91, 259, 300
Dielectric 200–1, 420
Diesel 121, 159, 263, 267, 276, 279, 300
auxilliary power unit (APU) 305
engines 347, 351, 387
fuels 266, 268, 304, 417, 420
low‐sulfur 277, 386
oil 323
synthetic 273, 323
systems 354
Diethyl sulfide 272
Differential pressure 116, 147, 153, 221
control valve 124
Diffusion 65, 115, 142, 427
barrier 302
bonding 107
characteristics 293
coefficient 76, 146, 166, 420
limitations 220
molecular 294
Index
overpotential 419
rate 240
water 91–2, 100, 165
Digester gas 223
anaerobic 417
Digestion 169, 273, 318
anaerobic 225, 272–3, 321
microbial 320–1
Digital storage oscilloscope 65
Diode 58, 368–73, 375, 420
Diphenyl isophthalate 78
Direct anodic oxidation 162
Direct borohydride fuel cell (DBFC)
17, 179–81
Direct carbon fuel cells (DCFC) 6,
17, 233–4
Direct combustion 272
Direct current (DC) 23, 368, 376, 378,
387, 391, 399, 415, 420, 426, 430
DC/AC conversion 125, 371, 378,
381, 384, 396
inverter 381
output 125, 365
power 205, 312, 351
regulators 366
voltage 35, 366, 370, 372
Direct ethanol fuel cell (DEFC) 158,
169–73, 182, 184
acidic 170
platinum‐free 172
Direct ethylene glycol fuel fells (DEGFC)
158, 174, 176
Direct formic acid fuel cell (DFAFC)
158, 176, 178
Direct fuel cell (DFC) 218, 227–9, 231–2
Direct internal reforming (DIR) 218–19,
282–3, 420, 423, 426
Direct liquid fuel cell (DLFC) 157–9, 161,
163, 165, 167, 169, 171–9, 181–3
Direct methanol fuel cells (DMFC) 17,
51, 67, 77, 157–8, 160, 162–3, 167,
183–5
anode 161–3
mixed reactant 115
passive 165
stacks 182
system 164–5, 182–3
Direct oxidation of hydrocarbons 284
Direct propanol fuel cell (DPFC) 158,
173, 174
Discharge 20–3, 334, 336, 338, 416,
418, 421
atmospheric pressure gliding arc 289
corona 290
dielectric barrier 290
microwave 290
Disproportionation 280, 420
Dissociation 85, 161, 293, 334
constant 421
Distributed generator (DG) 321, 378
Distribution networks 308, 372–3
Dopants 80–1, 239–41, 294, 432
acceptor 239, 425
Doped ceria 241, 243, 245
cermet 245
Doped perovskite type oxide 241
lanthanum chromite 247
Doping 213, 240–1, 244–5, 247
Dosing pump 227
Double‐layer 58–60, 65, 421
Dreamcar project 183–4
Drive‐train 35, 125, 127, 421, 431
Dry air 94, 100–1, 407
Drying 95, 172, 212, 433, 436
effect 94
rapid 105
Dry reforming 279
Dupont Corporation 69, 74, 428
Dye‐sensitized solar cell 313–14
Dynamic braking 381
e
Eaton supercharger 359–60
Edge connections 113, 149
Edge seals 14–15, 110
plastic 109
Edison, Thomas 231
Electronic conductivity 52, 236, 238–40,
243–7
Efficiency 25, 32–3, 35–6, 187–9,
312–13, 352–4, 356–7, 360–1, 391,
408–9, 411–13
cold gas 318
electrical 259, 282, 291, 296
443
444
Index
Efficiency (cont’d )
energy‐conversion 304
engine 387
fuel‐cell 35–6
isentropic 355–6, 361
limit 34
mechanical 356, 362
photon 313
solar 320
tank‐to‐wheels (TTW) 347
thermodynamic 33
well‐to‐wheels (WTW) 387
Electrical permittivity 59
Electric vehicles 20, 35, 125, 330, 350,
381, 395, 398, 421, 430
Electrocatalysis 171, 184
Electrocatalyst 4, 11, 76, 247, 306,
320, 421
Electrochemical impedance
spectroscopy (EIS) 61–2, 64–5,
68, 191, 421, 428
Electrochemical oxidation 170, 180,
209, 292
Electrochemical vapour deposition (EVD)
240, 249, 256
Electrode 10–16, 18–22, 45–52, 57–62,
88–92, 147–52, 431–6
area 10, 166, 212
catalysts 115, 152, 189
counter 60–1, 313, 419, 430, 436
degradation 85
kinetics 51, 421
photosensitive 430
potentials 85, 433
reactions 8–10, 21, 23, 47, 52, 81, 90,
136, 160, 162, 170, 423, 428, 436
rolled 147–8
titania 313
Electrolyser 44, 130, 261, 305–10, 312,
323, 385, 389, 422
high‐temperature 312
Electrolysis 2–4, 264–5, 305–8, 323, 333,
385–7, 416, 418, 422
Electrolyte 2–4, 6–15, 52–5,
89–93, 196–9, 207–16, 218–21,
234–46
Electromagnetic force 5
Electron 7–8, 29–32, 52, 58–9, 160–3,
170–1, 180, 312–13, 420–2, 424–6
donors 320
free 290
micrograph 82, 422
moles of 29, 207, 405
pairs 161–2
Electro‐osmotic drag 77, 92, 100, 142,
174, 422
Electro‐oxidation 173
Electrospray processes 84
Elenco 139–40
Embrittlement 425
Emissions 25, 184, 264, 276, 287, 387–8
reduction 383
Energy 7–9, 157–9, 348–51, 379–83,
387–9, 422–4, 431–3
activation 9, 415
density 182, 324, 349, 408, 422
Energy for You (EFOY) 157
Enthalpy 27–8, 35, 193, 296, 334, 345,
401, 408, 422–4
Entrained‐bed gasifiers 269
Entropy term 33, 312
Equilibrium 27, 48, 94, 193, 222, 277–8,
286–7, 393, 423
constant 222, 280, 421
thermodynamic 218, 280, 290, 390
Equivalent weight (EW) 74–5, 77, 166, 423
Ethanol 18, 23, 61, 83, 157, 159,
169–77, 264, 272, 274, 323, 325,
386–7, 417, 423
Ethylene glycol 142, 157, 174–6
Eutectic mixtures 211, 214, 423
Exchange‐current density 47–9, 51, 54,
57, 82, 117–18, 153, 394, 423, 434
Exergy 27, 193–4, 423
Expansion coefficient 244, 247
f
Fans 104, 106, 114, 120, 164, 306, 351
centrifugal 364
Faraday constant 29–30, 43, 166
Faraday, Michael 2–3, 5
Farads 59, 417
Fermentation 169, 272, 274, 320, 423
Ferritic steels 248
Index
Filling stations 308–9, 332–3, 385, 387
Fisher–Tropsch reactors 386
Flame traps 330
Flammability 168, 264
Flooding 73, 78, 90, 92, 102, 148, 164
Flow batteries 20–1, 131, 182, 423, 431
Flow‐field 12, 89, 102, 107, 109, 111–14
anode gas 215
interdigitated 11, 100, 103, 111
Flowsheet 385, 390–1
Fluorine‐containing polymers 146
Fluorite structure 240, 245
Formaldehyde 162, 168–9, 175
Formic acid 18, 157, 162, 169, 175, 177–8
Formic acid fuel cells (FAFC) 158, 176
fossil fuels 5, 26, 33, 232, 263, 265–8,
306, 416, 424, 434
Four‐quadrant inverter 383
Free energy 27–30, 37, 41, 117, 143, 160,
193–4, 222, 237, 404, 424, 428
barrier 418
change in molar Gibbs 401–4
of formation 28, 43
standard 28
Frequency response analysers (FRAs) 62
Fuel cartridge 123
Fuel cell 1–11, 13–21, 23–8, 30–41,
54–60, 91–8, 187–90, 378–85, 389–99
bio‐electrochemical 416, 428
borohydride 144, 178–9, 180, 342
circular 136
compact 70
direct acid 176
direct alcohol 171, 174–5
direct carbon 6, 17, 231, 233
mixed reactant 114
mobile 169, 277
portable applications 184
regenerative 132–3, 432, 435
stationary 265, 298
systems 192–5, 273–5, 278–9, 289,
291, 295–6, 351, 370–2, 378, 384,
387–9, 394–5, 397–8
Fuel cell vehicles (FCVs) 155, 157, 263,
267, 302, 304, 308, 323–4, 329–30,
332–3, 344, 347, 384–95
Fuel infrastructures 191
Fuel pretreatment 223
Fuel processing 76, 120, 188, 208, 263,
275, 281, 290–1, 321, 384, 389
practical 295, 304
Fuel utilization 39, 41, 177, 188–9,
191–2, 249, 391, 408
Fugacity 36, 415, 424, 434
Fullerenes 345
Functionalization 86–7, 145–6
g
Gadolinium 241, 294
Gadolinium‐doped ceria (GDC) 241
Gadolinium strontium cobaltite 247
Gallium 241
Gallium arsenide 313
Gas adsorption isotherms 416
Gas battery 5
Gas crossover 116, 214, 221
Gas density 357
Gas‐diffusion electrodes 198, 205
modern 155
Gas‐diffusion layer (GDL) 82, 89, 112,
114, 144
macroporous 164
Gas engines 274
Gas hydrates 268
Gasification 269, 273, 317, 424
Gas manifold 227
Gas micro‐turbines 398
Gasoline 157, 159, 168–9, 263–4,
266–7, 276, 279, 304–5, 347–8, 387,
420, 424, 429
Gas phase transport 220
Gas purification 293
Gas separation 287, 293–4, 343
Gas shift reaction 121, 189, 209, 275, 320,
434, 436
Gas‐tight seals 114
high‐temperature 249, 252
Gas‐to‐liquid 285
Gas turbines 222, 253, 258–9, 263,
398, 419
Gate turn‐off (GTO) 368
Gemini 6, 69–70
General Atomics 316
General Electric (GE) 69–70, 307
445
446
Index
General Motors (GM) 127, 304, 385–7
Gibbs 27–30, 36–7, 41, 117, 143, 193–4,
222, 237, 284, 418, 424, 428
energy 401
equation 28, 424
free energy of combustion 193
free energy of formation 28–9
function 401
Glass ceramics 248
Glow discharges 290
Glycolic acid 175
Gold 109, 180, 245, 291, 430
Grain boundary 424
Graphene 87, 91, 345
Graphite 12, 16, 107–9, 234, 345, 424
Greenhouse effect 419, 424–5
Greenhouse gases 263, 387, 419, 424–5
Grid 110, 372–3, 377–80, 399
electricity distribution 378
local 378
Grid‐connected systems 378
Grotthus mechanism 74
Grove 4–5, 7, 395
h
Half‐cell reaction 233, 425
Halogenated fluorocarbons 425
Harmonic distortion 373
Hat management 23, 238, 323, 349
Hazard identification (HAZID) 392
H‐bridge inverter circuit 372
Heat capacities 201, 355–6, 412–13
molar 402, 404
Heat engines 5, 32–3, 35, 187–8, 435
reversible 417
Heat‐exchangers 15, 19, 23, 155, 164,
188, 193–5, 205, 297–8, 302, 364,
390–1, 425
Heating and cooling curves 195
Heat losses 155, 193, 280, 286, 310, 354
Heat of reaction 284, 340, 423
Heat of vapourization 264, 409
Heat pumps 342
Heat sink 368
Heat storage 287, 342
Heat transfer area 193
Heat utilization 218, 227
Heliostats 287–8
Helmholtz 58
Heterocyclic compounds 339
Heteropolyacids 88
Hexis Ltd. 257
Higher heating value (HHV) 33, 274, 312,
327, 425
Hitachi 217, 228
Hole 15, 58, 110, 217, 239, 313, 360, 425
pairs 214
Honda 71, 127, 313
Honeywell 301
Hopping 74
vacancy 239
Horizon Fuel Cells 106, 116, 122–3,
182, 337
Hot Module 216, 226–7
Hotspot 303
Hot stream 194–6
Humidification 101–2, 107, 120, 412
cabin 142
Humidifier 23–4, 101, 116
Humidity 53, 92, 96–8, 100–1, 105, 110,
115, 120, 124, 197, 407
ratio 94
Hybrid cadmium cycle 316
Hybrid drive‐train 387
Hybrid electric vehicle (HEVs) 304, 383,
385, 387, 425, 429–32
Hybrid flow batteries 20
Hybrid sulfur cycle 316
Hybrid systems 156, 254, 256, 259–60,
380–82, 387
Hydrazine 142–3, 325, 341
Hydrides 122, 159, 310, 333–40, 343, 348–9
Hydrocarbon fuels 17, 168, 210, 245–6,
263, 284, 316, 409, 423, 429
synthetic liquid 273
Hydrocarbon oxidation 233, 245
direct 284–5
Hydrocarbon polymers 77
non‐fluorinated 78
Hydrocell 152
Hydrocracking 267
Hydrogen 7–10, 16–19, 25–30, 112–16,
124–7, 187–92, 263–5, 308–10,
317–21, 323–45, 347–9, 434–6
Index
alloys storage 180
carrier 340–1, 343
consumption 53, 138
distribution 340
economy 26, 304, 306, 425
electrode 48, 50, 422
embrittlement 293, 310, 330
energy 26, 28, 64, 72, 155–6, 158,
173, 212, 234, 289, 321, 345, 349,
350, 399, 408
fuel cell 27, 29, 31–2, 34–5, 37, 40, 43,
48, 50, 52, 182, 401, 403–4, 406
fuel‐cell 402
generation 264, 278, 289, 323, 350
in graphite nanofibre 344
ions 293, 433
oxidation 9, 82, 85, 158, 202, 394
production 23, 120, 235, 265, 277, 318,
320–1, 339, 341, 343
purification 292
safety 326
separation membranes 81
storage 131, 143, 323–5, 327, 329,
331–5, 337–9, 341–5, 347, 349–50
usage 405, 407
utilization 203
Hydrogenase 318–21
Hydrogenation 272, 305, 425
Hydrogenics 89, 115, 123–4, 308, 395
Hydrogenolysis 276
Hydrogen oxidation reaction (HOR) 8,
50, 82, 88, 203, 401
Hydrogen peroxide 83, 179
Hydrogen sulfide 204, 271, 274, 276, 419
Hydrolyse 179, 343
Hydrophilic 73
Hydrophobic 69, 73, 111, 164
Hydrostik hydrogen canister 337
Hydrous aluminium silicate minerals 436
HyFLEET 298
Hyundai 24, 127
i
IdaTech 301
Ignite 327, 333, 342
Ignition 326–7, 330
Ignition temperature 326–7
IMHEX bipolar plate 217
Imide‐forming amides 341
Impedance 62, 64, 190, 421, 425, 428
finite‐length Warburg 65
local 190
mass‐transport 64
Impedance spectroscopy 65, 421, 428
electrochemical 61, 65, 68, 428
Impregnation 90, 150, 198
Impurities 142, 189, 219, 234, 305, 337
Inconel 237, 425
Indirect internal reforming (IIR) 218,
282–3, 420, 423, 426
Inductance 373, 375, 380, 426
Inductor 368–72, 379, 426
Injection 308, 394
direct liquid water 102–3
moulding 108
Ink 83, 244
Inner Helmholtz plane (IHP) 59
Insulation 227, 302, 333
Insulators 114, 420, 432
electronic 11
Integrated gate bipolar transistor
(IGBT) 367–8, 370, 373, 426
Integrated Planar SOFC 257
Integrated reforming 283
Intelligent Energy 113–14, 123, 127, 324
Interconnect 139, 237, 247–50, 253,
419, 426
Intercoolers 116, 356
Interdigitated flow‐field design 102
Interfaces 52, 58, 62, 239, 252, 261,
392, 424
anode|electrolyte 245
cathode platinum–Nafion 64
electrochemical 261
electrode|electrolyte 44, 59, 415, 417–18
electrolyte|electrode 245
gas|electrocatalyst|electrolyte 435
solid|liquid 421
stable electrolyte|gas 210
three‐phase 198
Interfacial 418
Intermediary, important 319
Intermediate products 171, 176
partial oxidation to 175, 180
447
448
Index
Intermediate temperature solid oxide fuel
cell (IT‐SOFC) 237–8, 241, 244–8,
251, 257, 259
applications 245
electrolytes 243
temperatures 247
Internal combustion engine (ICE) 6–7, 116,
267, 329, 359, 384, 417, 425, 429–30
Internal combustion engines vehicle
(ICEV) 304
Internal currents 52–3
Internal manifolding 15–16, 216–18
Internal reforming 281
MCFC 275–6, 284
SOFC 393
Internal resistances 55, 59, 62, 282,
367, 426
Internal short‐circuit 140, 214, 426
Internal shunt currents 149
International Civil Aviation Organization’s
Dangerous Goods Panel 169
International Fuel Cells (IFC) 7, 18,
136–7, 199–200, 204–5, 298, 300, 397
International Union of Pure and Applied
Chemistry (IUPAC) 28, 422, 433
Inverter 127, 224, 366, 371–3, 376, 378,
380, 384, 391, 416, 426
circuits 380
grid‐connected 378
multilevel 375
pulse‐modulated 374
single‐phase 372–3, 377
Iodine 315–16
Ion conductivity 241
Ion conductors 239
fast 433
Ion‐exchange resins 201, 426
Ionic character 73, 80
Ionic compounds 421
Ionic conductivity 75, 145, 202,
240–1, 246
intrinsic 146
Ionic conductor 239, 435
Ionic liquids 79–80
room‐temperature 426
Ionic radius 80, 240–1
Ionomer 73, 76, 168, 426
Ion transport membrane (ITM) 301
Iridium 85, 129, 131, 430
Iron 86–7, 129, 150, 172, 205, 290, 317–18
austenitic 416
hydride 325, 335
oxides 93, 315–16, 390
receiver 232
Irreversible processes 31
Isentropic temperature 355
Ishikawajima‐Harima Heavy Industries
(IHI) 132, 228, 300
Islanding 378
Isotherm 336
pressure composition 335
Isothermal conditions 33, 312
ITM Power 308–11
ITM Syngas process 301
j
Jacques 232–3
Jet Propulsion Laboratory 182
Junction 10, 58, 245, 313, 367
triple‐point 4
k
Kawasaki 80, 332–3
Kerosenes 266, 284, 305, 420
Ketones 178
poly‐ether 146
KIMEX project 133
Kinetic energy 31, 193, 353, 355, 364,
431, 436
Kordesch, K. 135, 148–9, 156
Korea Electric Power Corporation
(KEPCO) 228
Kyocera Corporation 255–6, 396
l
Lanthanum 180, 241, 243, 246
chromite 237, 248
gallate 241
lanthanum strontium manganite
(LSGM) 233–4, 236, 238,
247–8, 255
molybdenum oxide 243
niobate 81
strontium cobaltite 247
Index
tungstate 81
vanadate 81
Lanthanum strontium cobaltite ferrite
(LSCF) 246
Laser ablation 345
Laves phase 334
Levelized cost 25
LG Electronics Inc. 258
Life‐cycle analysis 388–9
Lignite 416–17, 427, 434
Liquefied petroleum gas (LPG) 266, 268,
272, 324, 427
Liquid air 333
Liquid hydrocarbon 120, 269, 271–2,
305, 423, 434
fuels 347
Liquid hydrogen 155, 324, 327, 331, 333,
335, 348, 385–7
cryogenic 347
Lithium 19, 211–12
aluminate 207
amide 341
hydride 325
Lithium‐ion batteries 126, 130, 349
Load‐levelling 310
Losses 31, 34–5, 40–1, 43, 45–6, 49,
52–4, 56, 98, 100, 119–20, 210,
212–14, 370–1, 393
Lower heating value (LHV) 33, 35–6,
124, 136, 159, 169, 189, 229, 235, 296
Lquefied natural gas (LNG) 271
Luggin capillary 60, 427
Lurgi slagging gasifier 270, 284
Lysholm compressor 352–3, 357–8,
360, 411
m
Macrocyclics 86–7, 150
Magnesium 20, 241, 247, 337
Magnetic field 426, 432
induced 368
Manganese 148, 213, 416
Manganese dioxide 31, 150
Manganites 247, 315
strontium‐doped lanthanum 236
Manifolds 14–15, 114, 201, 216–17
Mass balances 195
Mass flow 407
factor 357–8, 411–12
non‐dimensional 357
Mass storage efficiency 329
Mass transfer 64, 286, 419
Mass transport 56, 65–7, 117, 131, 244,
393, 421, 427
Materials handling applications 183
Matrix 141–2, 152, 155, 197, 210,
212–13, 215–16
ceramic 207, 211–13, 221
electrolyte support 216
oxygen ion‐conducting 173
solid 210, 271
McDermott International 257
MC Power 228, 398
Mean time between failures 24
Mechanical energy 385, 421
Mechanical integrity 108, 317, 330, 344
Mechanical stability 213, 240–1, 245
Mechanical strength 197, 212, 416
Membrane 21–2, 74–8, 90–3, 103, 129,
145–6, 165–6, 180–2, 292–5
acid‐doped 78
anion‐conducting 181
bilayer 168
cation 180
cation‐conducting 181
cation‐exchange 426
ceramic 293
composite 172, 419
electrode 100
electrolyte 92, 96
hybrid 185
hydrated 165
ion‐conducting 181
ion‐exchange 426–7, 431
microporous 294
nanocomposite 77
non‐porous 293–4
performance 293
permeable 180–1
planar 301
proton‐conducting 145, 170
reactors 287, 295, 301
tubular 302
WGS reactors 293
449
450
Index
Membrane electrode assembly
(MEA) 12, 69, 83–4, 91, 93, 100,
109, 111–14, 165
Mercedes B‐class F‐CELL 125
Mercedes‐Benz buses 126
Metal alloys 334–5, 337, 415
Metal bipolar plates 107, 109–10, 129,
237, 251, 306
Metal dusting 281
Metal foam 102, 109–10
Metal hydrides 158, 326, 335–6, 338, 347
inorganic 341
practical 343
rare earth 326
simple 342
Metal interconnects 237
Metal membranes 294
Metal oxide 20, 293, 315
framework 346
proton‐conducting 294
Metal–oxide–semiconductor field]
effect transistor (MOSFET)
367–8, 370
Methanal 162
Methanation catalyst 291, 292, 308
Methane 208, 210, 218, 268, 270, 273,
275, 277–81, 284–6, 289, 292, 316–17,
320–1, 326–7, 416–17, 419
conversion 218, 282
electro‐oxidation 64
entrapped 330
hydrates 268
reformed 191
Methanex 168
Methanol 17–18, 121–2, 157–63,
165–70, 173–9, 182–4, 264–5, 267,
274, 279–80, 300, 303–5, 347–8,
386, 389
crossover 77, 167–8
decomposition 343
dehydrogenation 167
electro‐oxidation 161, 177
fuel cell 32
on‐board 347
oxidation 158, 161–3, 167, 170, 172
production 168
reformed 280, 292
reformer 291
safety and storage 168–9
sensor 164
Methylcyclohexane 339
Methyldimethoxysilane 77
Microbial fuel cell 23, 416, 428
Microchannel reactor (MCR) 301
Microcontrollers 130
Microcracks 241
Microdomains 74
Microelectrodes 64
Micro‐electromechanical system (MEMS)
113, 428
Microorganisms 417
living 428
Micropores 198, 294
Microporous 164, 258, 427
polyolefin membrane 22
separators 306
Microprocessor 129, 377
Microscope
cross‐section 302
optical 422
Microstructure 132, 244
Migration 3, 197–8, 216, 248, 295
potential‐driven 210
simple 74
Military bases 187
Millennium Cell 344
Mini‐pak 122, 123, 337
Ministry of Economy, Trade and Industry
(METI) 228
Mischmetal 180, 336
Mitochondria 169
Mitsubishi Electric 204–6, 228, 301
Mitsubishi Heavy Industries 251,
256–8
Mitsubishi Hitachi Power Systems
(MHPS) 255
Mitsui Engineering 257
Mixed reactant fuel cell (MRFC)
114–15
Mixed reforming 279
Mobile applications 17, 304
Mobile homes 182–3
Mobile phones 122, 182, 381
charging 71, 100
Index
Model 51, 54, 70, 99, 112, 119, 126, 192,
360, 421
dynamic 389, 392
economic 385
fuel‐cell electrodes 65
hybrid 383
mathematical 393
molecular 345
Module 6, 123–5, 130, 224, 306
multi‐tube 294
tubular membrane 302
Molarity 36, 428
Molecular biology 344
Molecular liquids 214
Molecular weight 264
Mole fraction 166, 428
Moles 29, 36, 96, 181, 194, 207, 279, 340,
390, 405, 408, 428, 434
Molten carbonate 17, 173, 210–11,
213–14, 222, 233, 235, 274
Molten carbonate fuel cell (MCFC) 207,
209, 211–13, 215, 217, 219, 221–3,
225, 233–4
anode 213
cathode 214
electrolyte 173, 211–12, 219
stack 215, 217, 219, 222–3, 225
systems 19, 195, 201, 207,
230–1, 361
Molten sodium 287
Molybdenum 109, 147, 245, 281, 425
Molybdenum nitrides 87
Monolayer 85
Monomers 78, 146, 428
α,β,β‐trifluorostyrene 77
functionalized 146
Monopolar 139, 151, 306, 428
Moon 6, 136
Moonlight Project 228
Motor 23, 116, 118, 361, 365, 381–2, 384,
387, 416, 424, 431
dedicated hub 388
electrical 365
electric‐drive 425
external 352
three‐phase 376
Motorbike manufacturer Suzuki 127
Motor controller 382
Motorola Labs 182
Multiwalled nanotubes 345
Municipal solid waste (MSW)
273, 428
Municipal wastewater 223
Murata Manufacturing 257
n
Nafion 64, 69, 72–5, 77–8, 145–6, 166,
172, 174, 176, 428
composite 173
solution 83
structure and characteristics 74
Nafion/silicon oxide 185
Nanocrystalline CoO photocatalyst 314
Nanocrystals 103
Nanomaterials 344
Nanomaterials for Solid State
Hydrogen 350
Nanotechnology 198, 344
Nanotubes 345
Naphtha 269, 278–9, 284, 323, 386
crude oil‐based 386
steam reforming of 278–9
Napier grass 272
National Aeronautics and Space
Administration (NASA) 69–70, 136,
331, 332
National Institute of Science and
Technology (NIST) 433
National Renewable Energy
Laboratory 292
Natural gas 189, 218–19, 225, 237,
256–9, 268–72, 274–5, 279–82,
297–8, 302–3, 308, 389–90, 419–20
composition 271
compressed 226, 386
grid 434
liquefied 271
liquids 430
pipelines 308, 310
steam reforming of 26, 168, 203, 272,
296, 312, 385, 434
synthetic natural gas (SNG) 273,
308, 434
vehicles 387
451
452
Index
NDC Power 172
Necar 125, 157
Neodymium 180, 243, 246
Nernst equation 36–41, 46, 55, 131, 207,
220, 222, 235, 237, 285, 295, 393, 428
Nernstian effect 153
Nernstian losses 46
Nernst relationship 189
Nernst voltage 253
n‐ethylcarbazole 339
Networks 310, 323, 373, 375, 421
controller area 130
gas transmission 308
linked 433
local 125
New Energy Development Organization
(NEDO) 228
Nickel 85–7, 147–50, 205, 208, 213–15,
219, 235–6, 238, 241, 243–6, 273, 276,
279–80, 285, 317
amorphous alloy membrane 293
anode 150, 173, 180, 213
catalyst 150, 284
cathode dissolution 214–15
cermet 284
crystallites 281
electrocatalyst 209
electrode 147
ions 214
matrix 255
mesh 148–9
porous anode 219
sintering 245, 279
Nickel nitrate 87
Nickel oxide (NiO) 153, 208, 210–11,
214–15, 244, 317
Nickel‐YSZ 243
Nicotinamide adenine dinucleotide
phosphate (NADP) 319
Niobium 109, 213, 293
Niquist plot 428
Nissan 71, 127, 304
Nitric acid 87
Nitrides 87, 132
Nitrogen 86–7, 94, 96–7, 154, 220, 271,
274, 295, 340–1
Nitrogen oxides 230, 263, 298, 416
Noble metals 109, 131, 209, 236, 419
incorporating 246
Non‐catalytic partial oxidation (NCPO)
302–3
Non‐conductive coatings 290
Non‐noble metal catalysts 158
Non‐porous
ceramic membrane 294
metal membrane 293
Non‐pyrophoric 291
Non‐retum flap 227
Non‐slagging 270
Non‐stoichiometric 80
Non‐sulfonated membranes 77
Non‐volatile metal oxide cycles 316
Notice of Market Opportunities
(NOMO) 7
Nucleophiles 146
Nyquist plot 62–4, 428
o
Odourants 272, 275–6
Ohmic loss 46, 54–5, 59, 65–8, 152, 197,
202, 212, 238, 240, 242, 245, 253–5,
257, 428–9
internal 190
Oil shales 268
Open‐circuit voltage (OCV) 27, 29–35,
37, 39–41, 43–6, 49, 51–3, 66, 117,
153–4, 159, 187, 189–92, 237, 428–9
Operating voltage 24, 35, 41, 57, 59, 212,
222, 312, 391
Orbiter 140, 150, 153
Organic frameworks 86, 346
Original equipment manufacturers
(OEMs) 257, 360, 395, 429
Osmotic drag 77, 78
Outer Helmholtz plane (OHP) 59
Overpotential 40, 44, 47–8, 50–1, 57, 59,
64–5, 99, 150, 201, 204, 307, 312,
429–30, 434
Oxide ion 240
Oxygen carriers 317–18
Oxygen concentrators 295
Oxygen evolution reaction (OER)
131, 315
Oxygen‐ion‐conducting membranes 295
Index
Oxygen reduction reaction (ORR)
87, 131, 170, 173, 176, 230
Oxygen transport 164
Oxygen usage 406–7
Oxygen utilization 190, 192
82, 85,
p
Palladium 51, 168, 172, 174, 177, 246,
292–93, 302, 430
Parabolic dish 288
Paraffins 267
branched 266
Para‐hydrogen 333
Parasitic losses 103, 228, 365
Parasitic power load 292
Partial oxidation (POX) 173, 175, 180,
270, 285–6, 416, 429
and autothermal reforming 285
reactors 303
Partial pressure 36, 38–9, 55, 94–6,
131, 189, 192, 202, 207, 214–15,
221, 241, 247–8
Passive systems 165
PCT curve for hydrogen absorption and
desorption 335
Perfluorinated sulfonic acid (PFSA)
membrane 72
Performance charts 356, 358–60,
363, 413
Permeation 293
controlled 148
Perovskite 150, 241, 246–7, 429
complex 80
cubic 241–2
lanthanum cobaltite 247
structures 80, 241, 247
Petroleum 234, 266–8, 323, 331, 416,
420, 428, 430
crude 266
fractions 276, 279, 285
refining 429
PFSA‐type membrane materials,
structure of 75
Phase angle 62, 64, 379–80, 435
Phosphoric acid 17, 69, 78, 80, 196–200,
210, 235
concentrated 196
Phosphoric acid fuel cell (PAFC) 17–19,
38, 69–70, 136, 150, 187–9, 191–3,
195–9, 201–6, 208–10, 220, 274–5,
297, 340–1
catalysts 198–9
degradation 205
development of 204–5
high‐temperature PEMFCs 280
stacks 189, 197, 199, 201, 275
stationary power 19
Phosphorous 243
Photobiological hydrogen
production 430
Photoconductivity 396
Photoelectrochemical cell solar 430
Photoelectrode 313
Photoelectrolysis 323
Photolysis 313, 430
electrochemical 313
Photosensors 58
Photosynthesis 313, 318, 417
Photosynthesis‐algae 321
Photovoltaic (PV) 26, 308, 310,
372, 391
array 378
cells 312–13, 389, 395
Phthalocyanines 86, 205
Physisorption 344–5, 415, 418
Pinch point 195–6
Planar configurations 248, 251
Planar SOFCs 238, 252, 254, 256–7
commercialization of 252, 256
single 252
Planar stack configurations 249
Plasma 84, 289
non‐thermal 289–90
spraying 249, 252
Plate reformers 283, 300
Platinum 18, 70, 81–5, 88, 148, 150,
161–3, 167, 171–2, 174, 177, 198,
204–5, 246–7, 291
alloy 85, 163, 175, 205
anode catalyst 103, 150
bulk 84
catalyst 18, 50, 61, 76, 84, 87, 121, 142,
147, 149, 165, 176, 198, 285, 339
dispersing 198
453
454
Index
Platinum (cont’d )
dissolution of 88, 129
loading 82, 175
particles 77, 84, 88–9, 199
water‐soluble 82
Plug Power 123, 300
Polarization 45, 429–30
Polyaniline (PANI) 87–8
Polybenzimidazole (PBI) 77–8, 146, 168
acid‐doped 77
alkali‐doped 170
Polybenzoxazine (PBOA) 78
Polycrystalline 396
Polyethylene 69, 72, 108, 310, 339, 344
Polyethylene oxide 146
Polyethylene terephthalate (PET) 174
Poly‐fluorinated sulfonic acid (PFSA)
72–8, 92, 159
ionomers 75
membranes 76, 159, 165, 176–8
Polymer 8, 11, 69, 75, 77–80, 87, 108–9,
123, 169, 200, 293, 341, 419, 426
acid‐doped 78
anion‐exchange 145
anionic 137
films 146
hydrocarbon 146
membranes 69, 92, 129
non‐fluorinated 77
perfluorinated 72
permselective 344
Polymethyl methacrylate 344
Polyphenolic resin 78
Polyphenylene sulfide 77, 108
Polypropylene 108
Polypyrrole (Ppy) 87
Polysulfone 145
Polysulfonic acid 172
membranes 115
Polytetrafluoroethylene (PTFE)
69, 72–3, 83, 90, 147–50, 164,
197–8, 210
Porphyrins 86, 150, 267
Potassium 19, 207, 281
Potassium hydride 325
Potassium hydroxide 136, 305
Potentiostat 61, 430, 436
Power capacity 22
Power conditioner 226, 351, 430
Power density 24–5, 70–1, 113, 115, 128,
159, 163, 172–3, 182, 204, 230, 251,
252, 256–7, 383, 431
Power distribution 375
Power electronics 24, 365–7, 384
Power factor 380, 431
Power factor correction (PFC) 378, 380
Power generation 25, 142, 273, 312,
395, 421
dispersed 265
distributed 421, 425
Power‐to‐gas (P2G) 308
Praseodymium 243, 246
Praxair 302
Preferential oxidation (PROX) 291, 390,
431–2
Preheaters 227, 259
Preheat 193, 208, 249, 296, 303, 356
combustion air 195
recycle gas 195
Pre‐ignition 267
Pre‐reforming 280–2, 390
Pressure 27–8, 36, 56, 94–6, 115–20,
126–7, 153–5, 201–2, 221, 268–70,
278–80, 323–4, 331, 334–5, 358–9
standard 30, 34, 36–7, 41, 402
Pressure lift 361, 364–5
Pressure ratio 119, 352, 357–9, 363, 411
Pressure swing adsorption (PSA) 292–3,
295, 431
Process flow diagram (PFD) 196, 298,
354, 389, 391
schematic 123
Process flowsheet 390
Process integration 188, 296, 310
Propane 223, 255, 268, 270–2, 274–5,
285, 289, 326–227, 340, 423, 427
Propanol 157–8, 173
Proton conductivity 74, 76–8, 81, 146,
168, 172, 176, 294
Proton Energy Systems Inc. 132
Protonex 255
Proton‐exchange membrane (PEM) 6, 8,
17, 69, 71, 73, 83–85, 91–93, 105, 144,
157–58, 166, 306–7, 427–8, 431
Index
catalysts 61, 198, 205, 292
cells 311
conductive 168
crossover 85
electrode 89
electrolyser 307, 310, 312
electrolytes 103, 160, 165, 170, 174
low‐temperature 82, 280–1
stacks 69–70, 88, 94, 100, 102, 105,
107, 112, 125–6, 130, 152, 259, 292,
304, 306
Proton‐exchange membrane fuel cell
(PEMFC) 17–18, 69–73, 75–85,
87–91, 93–9, 103–5, 107–9, 115–17,
119–21, 131–3, 135–7, 144–5,
157–67, 189–94, 275–7
Proton OnSite 308
Proton transfer 174
Pulse‐width modulation (PWM) 373,
375, 378
Pumping 197, 351
electrochemical 197
water 352
Pumps 13, 19, 21, 23–4, 102, 124, 138–9,
164, 231–2, 351, 358, 362, 364–5
PURASPECTM, 276
Purification 135
Pyrochlores 81
Pyrolysis 82, 245, 272, 289, 419, 424
anaerobic 273
Pyrophoric 291, 342
q
Quality standards 378
Quantum 313, 394, 420, 431
Quantum Leap Technology 397
Quaternary ammonium (QA) 145
r
Radial turbine 361–3
Radiation 104, 298–9, 313
electromagnetic 9
infrared 424–5
Radiation heat loss 238
Radiators 23, 76, 174
Radicals 129, 421
Radio‐frequency (RF) 290
Rail locomotives 267
Rand, D. A. J. 111, 156, 163, 175, 181,
187, 234, 277, 287, 321, 323, 350–1,
387, 399
Raney metals 147, 149–50
Rate constant 434
Reaction intermediates 176
adsorbed 171, 177
carbonyl 170
Reaction kinetics 115, 343
Reaction mechanisms 61, 170–1,
286, 421
Reactive power/kVAR 379
Real power 379
Rechargeable batteries 19–20, 130, 182,
310, 337, 348, 380–3, 423, 429, 431
Rectification 368
Recuperator 193
Redflow Pty Ltd. 22
Redox couples 423
Redox flow batteries (RFBs) 20, 130–1
Reference states 28, 194, 401, 423
Reformate 283, 292, 305, 431, 434
Reformed hydrocarbons 120–1, 408
Reformer 56, 116, 159, 189, 193, 205, 276,
278–9, 283, 289–90, 296, 298–300,
385, 390, 392
autothermal 295
catalyst 302
external 19, 219
industrial 296
on‐board 264, 347
plates 283
product gas 276, 292, 296
reactor 193, 281, 283, 291–2, 298–9
Reforming reaction 120–1, 188–9,
210, 219, 221, 277–8, 282–6,
299–300
Regenerative braking 382
Regenerative fuel cell systems 131
Regulator circuit 369–70
Relative humidity (RH) 75, 92, 94–5,
97–101, 105, 163, 291, 432
Renewable energy
sources 155, 306, 310, 351, 387
storage 324
Reversible adiabatic expansion 415
455
456
Index
Reversible metal hydrides 326, 333, 338
Reversible processes 27, 31, 193, 424
Reversible voltage 312
Rolled carbon 148
Rolls‐Royce 237, 257–8
Roots compressors 352–3
Rotating disc electrode (RDE) 61, 85
Roughness 50–1
Round‐trip efficiency 131, 307, 432
Rupture discs 330–1
Ruthenium 85, 88, 131, 163, 287, 305,
313, 430
s
Safety 326
Samarium‐doped ceria (SDC) 241
Sankey diagram 124–5
Saturated vapour pressure 95
Screen 66, 257
printing 252
silver‐plated nickel 150
Scrubber 137
Sealing 89, 215–16, 218, 237–8, 252
gasket 106
Seals 11, 113, 143, 151, 216, 248, 251
Selectivity 172, 176
Selenides 87, 418
Self‐discharge 429
Self‐humidification 99–106
Semiconductors 58, 230, 396, 425, 428
crystal 239
gadolinium 313
nanomaterials 314
n‐type 214, 247, 312, 428
three‐terminal power 426
Sensors 326, 377, 392, 428
Separation membranes 294, 302
Separator 217, 432
Sequestration 317–18, 432
Serpentine 11, 110, 111, 424
Shift 189, 209, 222, 235, 259, 275, 286
catalyst 291
converter 297
reaction, reverse 280
Short‐circuit 152, 378, 426, 432
partial 11
Shuttle 17
Siemens 135, 139, 147, 153, 174,
254, 398
Siemens Westinghouse Power Corporation
(SWPC) 249–50, 258
Silane 325
Silent discharges 290
Silica 77, 103, 172, 243, 248, 267, 277,
294, 312, 396
membranes 173, 294
Silicon PV cell 396
Silver 1, 61, 147–50, 430
Simulated coal gas (SCG) 203
Single‐walled nanotubes 345
Sintering 88, 129, 213, 219, 243–4, 433
Sliding discharge reactor 289
Sludge 225, 273
sewage 273, 321
Slurry 212–13
organic 343
Smartphones 123, 337
Soda lime 137
Sodium 73, 181, 339, 342–3
Sodium borate 179
Sodium borohydride 82, 142–4, 157,
178–9, 265, 325, 342–3
fuel cells 181
Sodium bromide 21
Sodium carbonate 19, 136, 207, 212
Sodium dithionate 178
Sodium hydride 325
Sodium hydroxide 136, 343
Sodium sulfate 73
Sodium tetrahydroborate 342
Solar 26, 132, 264–5, 306, 308, 387
embodied 286
energy 319, 383
hydrogen production 321
panels 372, 380
thermal 287
Solid oxide fuel cell (SOFC) 17–19, 173,
188–90, 234–41, 243–5, 247–9,
251–5, 257–61, 281–4
auxiliary power 305
high‐temperature 173, 237
intermediate‐temperature 237
internal‐reforming 281
low‐temperature 238
Index
multistage 258
stacks 189, 252–3, 258, 282–384
tubular 251
Solid polymer fuel cell (SPFC) 69
Solid State Energy Conversion Alliance
(SECA) 254
Solid‐state hydrogen storage 350
Sorbent‐enhanced reforming 287
Space applications 120, 153
Spacecraft 139, 141–2, 324
manned 69
Spark‐ignition 386, 429
Speed factor 358, 363
rotational 357
Stack 11–12, 14–15, 103–4, 115–16,
118–20, 122–5, 128–231, 283,
391–2
area 136
components 213, 217
configuration 215
cooling 192, 200, 282
networking 231
Stainless steel 12, 16, 107, 109, 215,
217, 242
Standard temperature and pressure (STP),
conditions 28, 30, 45, 53, 202, 264,
357, 407, 408, 433–7
State‐of‐charge measurement 383
Stationary power applications 23, 70,
125, 390
Steam electrolyser 307
Steam electrolysis 422
Steam methane reforming (SMR)
188–9, 321
Steam reformer 23, 273, 290, 296–8
Steam reforming 188, 277
Steam turbines 6, 33, 187, 194, 258, 362,
384, 419
Step‐down converters 368, 370–1
Step‐up converters 371
Stoichiometric 124, 129, 326, 405,
407–9
Storage 5, 116, 127, 131, 159, 168–9,
233, 323–4, 328–9, 331, 340,
349–50, 417
composite tank 330
cryogenic 333
high‐pressure 333
tank 22, 127
temporary 417
underground 288
vessel 159, 169, 330, 349
Stored energy 337, 380, 431
accessible 422, 433
Strontium 241, 247
cerates 80
Strontium titanate 246
Sub‐bituminous coal 416–17, 427, 434
Submarines 147, 154
Sub‐stacks 201
Sulfides 87, 313, 418
Sulfonated fluoroethylene 73
Sulfonated polyether ether ketone
(SPEEK) 77, 168
Sulfonate groups 83
charged 177
Sulfonation 73
Sulfonic acid 73
benzyl 77
group 75, 144, 166
perfluorinated 72, 172
polystyrene 69
trifluorostyrene 77
Sulfur 88, 189, 198, 204, 213, 219, 235,
267–8, 271–2, 275–7, 285, 291, 305,
315–16, 319
compounds 189, 276
content 276, 420
organic 275–6
poisoning 204, 238
removal 23, 296, 304
tolerance 291
Sulfur dioxide 285, 416
Sulfuric acid 3, 145, 315, 433
concentrated 73
Sulfur oxides 230
Sulzer Hexis AG 398
Supercapacitor 383–4, 421, 427
Superchargers 351, 361
Supported electrolytes 242
Supported nickel catalysts 219, 277,
286, 305
Supported Pt 292
Support electrocatalysts 417
457
458
Index
Surface area 50, 59, 83, 198, 213, 243,
416, 433
active 84, 88
available active 199
effective 10
Surface coating 107
Surface concentration 245, 419
Surface properties 85, 293
Sweep rate 84
Switching regulator 368
Switching sequence 374, 377
Switch‐mode 368–9, 371
Synthesis gas (syngas) 225, 233, 272–4,
277, 285–6, 289, 295, 298–304
Synthetic diesel fuel 268
Synthetic natural gas (SNG) 273,
308, 434
t
Tafel 393
equation 46, 48–9, 417, 434
plots 47, 57, 434
Tank‐to‐wheels 387
Tantalum 205, 293
Tape calendaring 252
Tape‐casting 213
Target program 205
Tars 269
Teflon 150
Telecommunications 148
towers, remote 71, 123, 274, 389
Temperature 27–30, 33–7, 94–9, 152–5,
189–97, 201–4, 222–3, 241–5,
268–70, 278–82, 284–7, 342–3,
354–7, 401–4, 432–3
distribution 192, 217, 219
gradients 216, 418
standard conditions 28, 433
Terminals 30, 429, 435
Tetraazaannulene (TAA) 86, 205
Tetraethyl orthosilicate 294
Tetrafluoroethylene 72
Tetrahydrofuran 347
Tetrahydrothiophene (THT)
272, 276
Tetramethoxyphenylporphyrin
(TMPP) 86, 205
Tetramethyl orthosilicate 294
Tetraphenylporphyrin (TPP) 86, 205
Thermal cracking, of hydrocarbons 289
Thermal cycling 221, 252, 257
Thermal decomposition 342, 417, 424
controlled 234
Thermal durability 277
Thermal efficiency 188, 287, 419, 435
Thermal expansion coefficient 237, 239,
248, 435
Thermal plasmas 289–90
Thermochemical methods, of hydrogen
generation 264
Thermoneutral condition 286
Three‐phase boundary 4, 11, 83, 90, 244,
246, 435
Three‐phase inverters 376–7
Thyristor 367–8
Tile 211–12
Titanium 109, 163, 205, 248, 293, 335
Titanium carbide 132
Titanium dioxide 77, 103, 172,
313, 421
Titanium hydride 325
Titanium nitride 109
Tokyo Electric Power 204–5
Tokyo Gas Co., Ltd. 257, 302
Tolerance band 375
Toray paper 424
Tortuosity 76, 435
Transistors 58, 367, 370, 396
Transition metal 109, 205, 246,
293, 334
chalcogenide 167
macrocyclic compounds 86
nitrides 87
Transmission systems 271–2, 310
high‐voltage 378
Trifluorostyrene 77
Tube‐and‐shell configurations 294
Tube trailer 328
Tubular cells 237, 251–3, 255
flat 256
Turbine 25, 187–8, 274, 332, 351, 354,
357–8
axial 361
calculations 362
Index
performance chart 413
wind 309–10, 378, 389
Turbochargers 357, 359, 361
Twin‐screw compressor 352–3, 360
u
UltraBatteryTM 383
Ultracapacitor 421
Ultracell 113
UltraFuelCell 258
Unicellular 318–19
Uninterruptible power system (UPS) 125,
255, 389
Union Carbide 135, 139, 148
United Technologies Corporation
(UTC) 6–7, 19, 140, 199, 205–6,
223, 225, 298, 397
Unitized regenerative fuel cell
(URFC) 130, 133, 432, 435
Universal three‐phase inverter 377
Utilities 188, 272, 378, 426
Utilization 155, 194, 202–3, 220, 222,
258, 350, 435
UV region 313
UV water purifiers 123
v
Vacancies 240, 313, 425
oxygen‐ion 239
Valency 241, 423, 435
Vanadium 22, 109, 205, 241
redox battery 21, 131, 182
Vapour 101, 165, 169–70
metal chloride 240
Vapourization 33, 210, 264, 409,
425, 427
Vapour phase 339, 409
Vapour pressure 37, 96–8, 100–1, 153–4,
176, 197, 264
Volmer, see Butler–Volmer equation
Voltage 35–7, 39–41, 43–6, 50–6, 59–62,
65–7, 117–18, 129–30, 192, 366–8,
370–80, 417–18, 421–2, 428–30, 435
degradation rates 228
fuel‐cell 34, 44, 54
losses 30, 34, 45–6, 55–6, 67, 82, 119,
149–50, 160
open‐circuit cell 190–1
regulator 23, 366, 369, 371
reversible cell 160, 222
Voltammetry 420, 436
linear‐sweep 436
Voltammogram 61, 436
cyclic 61, 177, 178
Volt–amperes 379, 431, 435
reactive 379
Vulcan XC72, 82, 163
w
Waste 154, 165, 265, 351, 417
agricultural 321
household 273
municipal 273
organic 320–1
Water‐cooled PAFC stacks 199–200
Water cooling 105, 123, 125, 128,
201, 390
Water evaporation 94, 165, 409
Water management 70, 72, 92, 129, 132,
142, 165, 307, 312
Water production 92, 407–9
rate of 101, 165, 405, 408–9
Water splitting 313, 318, 343
Water vapour 94–5, 97, 103, 109, 116,
164, 189, 202, 263, 343, 407, 425
pressure 96, 101
Wavelengths 313, 319, 425
Well‐to‐wheels 304, 385, 387, 389
Westinghouse 237, 254–6, 398
Wet‐proofing agent 210
Wet‐seal 215–18, 222
Wind 265, 306, 308, 310, 387,
419, 432
Wind farms 231, 380
Window‐frame design 253
Wood 103, 341, 417
pulp 178
wastes 417
x
Xerogel 294, 433, 436
X‐ray 146, 436
X‐ray photoelectron spectroscopy
436
394,
459
460
Index
y
Yeasts 169, 423
Yttria 235, 239
stabilized zirconia 233, 238–41, 244–7,
255, 258, 302, 422
z
Zeolites 293–5, 346, 436
faujasite‐type 277
separation membranes 155
Zetek–Zevco technology 139
Zinc 1, 20, 22–3, 31, 51, 147–8, 241
Zinc–bromine battery 182
Zinc oxide 276–7, 280
Zirconia 212, 216, 235, 239–40,
245, 317
calcia‐stabilized 239
cubic 240
tetragonal 240
Zirconium phosphates 77, 142