You are on page 1of 909

Catalysis by

Ceria and Related


Materials
2nd Edition

P870_9781848169630_tp.indd 1 8/4/13 2:17 PM


CATALYTIC SCIENCE SERIES

Series Editor: Graham J. Hutchings (Cardiff University)

Published

Vol. 1 Environmental Catalysis


edited by F. J. J. G. Janssen and R. A. van Santen

Vol. 2 Catalysis by Ceria and Related Materials


edited by A. Trovarelli

Vol. 3 Zeolites for Cleaner Technologies


edited by M. Guisnet and J.-P. Gilson

Vol. 4 Isotopes in Heterogeneous Catalysis


edited by Justin S. J. Hargreaves, S. D. Jackson and G. Webb

Vol. 5 Supported Metals in Catalysis


edited by J. A. Anderson and M. F. García

Vol. 6 Catalysis by Gold


edited by G. C. Bond, C. Louis and D. T. Thompson

Vol. 7 Combinatorial Development of Solid Catalytic Materials:


Design of High-Throughput Experiments, Data Analysis,
Data Mining
edited by M. Baerns and M. HoleÁ a

Vol. 8 Petrochemical Economics: Technology Selection in a Carbon Constrained World


by D. Seddon

Vol. 9 Deactivation and Regeneration of Zeolite Catalysts


edited by M. Guisnet and F. R. Ribeiro

Vol. 10 Concepts in Syngas Manufacture


by J. Rostrup-Nielsen and L. J. Christiansen

Vol. 11 Supported Metals in Catalysis (2nd Edition)


by J. A. Anderson

Vol. 12 Catalysis by Ceria and Related Materials (2nd Edition)


edited by A. Trovarelli and P. Fornasiero

Catherine - Catalysis by Ceria.pmd 1 4/4/2013, 3:26 PM


CATALYTIC SCIENCE SERIES — VOL. 12
Series Editor: Graham J. Hutchings

Catalysis by
Ceria and Related
Materials
2nd Edition

edited by

Alessandro Trovarelli
Università di Udine, Italy

Paolo Fornasiero
Università di Trieste, Italy

Imperial College Press


ICP

P870_9781848169630_tp.indd 2 8/4/13 2:17 PM


Published by
Imperial College Press
57 Shelton Street
Covent Garden
London WC2H 9HE

Distributed by
World Scientific Publishing Co. Pte. Ltd.
5 Toh Tuck Link, Singapore 596224
USA office: 27 Warren Street, Suite 401-402, Hackensack, NJ 07601
UK office: 57 Shelton Street, Covent Garden, London WC2H 9HE

British Library Cataloguing-in-Publication Data


A catalogue record for this book is available from the British Library.

Catalytic Science Series — Vol. 12


CATALYSIS BY CERIA AND RELATED MATERIALS
Second Edition
Copyright © 2013 by Imperial College Press
All rights reserved. This book, or parts thereof, may not be reproduced in any form or by any means,
electronic or mechanical, including photocopying, recording or any information storage and retrieval
system now known or to be invented, without written permission from the Publisher.

For photocopying of material in this volume, please pay a copying fee through the Copyright
Clearance Center, Inc., 222 Rosewood Drive, Danvers, MA 01923, USA. In this case permission to
photocopy is not required from the publisher.

ISBN 978-1-84816-963-0

Typeset by Stallion Press


Email: enquiries@stallionpress.com

Printed in Singapore

Catherine - Catalysis by Ceria.pmd 2 4/4/2013, 3:26 PM


b1469 Catalysis by Ceria and Related Materials

PREFACE

Ceria-based oxides have gained a major position in the list of non-


innocent supports, active co-catalysts and advanced materials, mainly
for environmental applications and for energy conversion systems.
The tremendous scientific interest is demonstrated by the fact that
in every year since 2006 over a thousand publications on related top-
ics are printed, a doubling of the already remarkable average num-
ber of manuscripts published every year in the previous five years.
Using the keywords “ceria/CeO2/cerium dioxide” and “cataly*” a
literature search returned almost 6,000 papers published since 1990,
many of which are from the last decade. The new century has con-
firmed the unique and irreplaceable role of ceria and ceria–zirconia
in advanced car converters for exhaust gas pollution control. Parallel
to this worldwide application, these materials have a bright future
in many new sectors, including as catalysts in hydrogen production
and purification and in fuel cells (either as electrolytes or catalytic
materials) and even as photocatalysts. One remarkable drawback
of the increasing popularity of ceria and related materials is their
availability, which is often dictated more by strategic and geopolitical
issues than technical issues. Cerium is the most abundant of the so-
called rare-earth elements, and is more abundant than copper or
nickel. Mining strategies, advances in extraction and separation
processes of cerium and its recovery from spent materials in con-
junction with strategic and geopolitical aspects, determine its actual
availability and possibilities for future applications. And this is true
of the entire market for rare-earth elements.
In this context, it is not surprising that the success achieved by
the volume Catalysis by Ceria and Related Materials, published in 2002,

b1469_FM.indd v 4/8/2013 12:43:14 PM


b1469 Catalysis by Ceria and Related Materials

vi Preface

prompted the publisher to suggest a second edition to review the


current work in this area. Based on the vastness of the subject, more
appropriate for an encyclopaedia rather than a single volume, we
decided to prepare a book with a good balance between conven-
tional applications and the study of ceria-based materials, and a
fundamental overview of “traditional” systems and strong view of the
emerging trends. Particularly important in our view are the novel
properties that nanostructured ceria has with respect to the bulk
material.
The book starts with a detailed and fundamental structural char-
acterization of the main ceria-based materials (Chapter 1), followed
by a careful overview of recent achievements in HR-TEM characteri-
zation (Chapter 2). Chapter 3 gives an insight into the oxygen
storage and release capacity of both model and commercial ceria-
containing three-way catalysts, which is one of the key properties in
their successful application. A discussion of the role of ceria-based
materials in the activation and conversion of nitrogen oxides
(Chapter 4) closes the first part of the book.
The second section of the book considers the peculiar aspects of
CeO2-based nanosystems, starting with an introduction to the analy-
sis of their structural complexity by means of atomistic modelling
(Chapter 5). An overview of synthetic strategies and characteriza-
tion techniques to obtain 2D and 3D nanoarchitectures is given in
Chapter 6, while the following chapter deals with recent advances in
the use of metal core/ceria shell systems to gain novel catalytic and
morphological properties.
The third part of the book (Chapters 8–14) is on the use of
ceria-based materials in emerging and commercial applications.
Chapter 8 discusses synthetic strategies for the preparation of ceria-
based catalysts for oxidation and combustion. Chapters 9 and 10
carefully investigate hydrogen purification reactions, namely the
water gas shift reaction (WGSR) and preferential oxidation (PrOx).
Chapter 11 discusses the challenges in the use of ceria-containing
catalysts for diesel soot combustion, while Chapter 12 investigates
the role of these materials in solid oxide fuel cells (SOFCs) and in
oxygen separation membranes. Finally, an overview of the potential

b1469_FM.indd vi 4/8/2013 12:43:14 PM


b1469 Catalysis by Ceria and Related Materials

Preface vii

application of ceria-based catalysts in the transformation of oxygen-


ated compounds in biomass into added-value products is presented.
The book ends with a panorama of the application of ceria-based
catalysts for air pollution abatement.
We thank all the contributors to the volume and we hope that
the book can stimulate further step changes in the sector while con-
tinuing to be a basic textbook for those who approach these interest-
ing and surprising materials.

The Editors
Alessandro Trovarelli
Paolo Fornasiero

b1469_FM.indd vii 4/8/2013 12:43:14 PM


b1469 Catalysis by Ceria and Related Materials

This page intentionally left blank

b1469_FM.indd viii 4/8/2013 12:43:14 PM


b1469 Catalysis by Ceria and Related Materials

CONTENTS

Preface v

Chapter 1 Crystal and Electronic Structures, Structural


Disorder, Phase Transformation, and Phase
Diagram of Ceria–Zirconia and Ceria-Based
Materials 1
Masatomo Yashima
1.1 Phase Diagram of the Ce–O System and Structural
Properties of Fluorite-Type CeO2 1
1.2 Phases and Structure of Zirconia and Ceria–Zirconia
Solid Solutions 4
1.2.1 Phase transition and crystal structure of zirconia 5
1.2.2 Crystal structure of monoclinic CexZr1−xO2
and tetragonal-monoclinic phase transformation
in the CeO2−ZrO2 system 7
1.2.3 Phase identification and crystal structure
of tetragonal phase in the compositionally
homogeneous CexZr1−xO2 − t, t ′, and t ″ forms 11
1.3 Phase Diagram and Partitionless Transformation
in the CeO2–ZrO2 System 15
1.3.1 Experimental and calculated equilibrium
phase diagram of the CeO2−ZrO2 system 15
1.3.2 Temperature-induced structural change
of tetragonal CexZr1−xO2 and partitionless
tetragonal–cubic phase transformation
in compositionally homogeneous CexZr1−xO2
materials 17

ix

b1469_FM.indd ix 4/8/2013 12:43:14 PM


b1469 Catalysis by Ceria and Related Materials

x Contents

1.3.3 Thermodynamics of the formation of t′-CexZr1−xO2


and nano-sized t′-CexZr1−xO2 22
1.4 Electronic Structure and Chemical Bonding of
Zirconia, Ceria, and Ceria–Zirconia Solid Solutions 24
1.4.1 Optimized structure and valence-electron density
of tetragonal zirconia and ceria 24
1.4.2 Optimized structure and valence-electron density
of tetragonal ceria–zirconia solid solutions 26
1.5 Structural Disorder and Diffusion Pathways of Oxide
Ions in Ceria–Zirconia and Ceria–Rare-Earth Oxide
Solid Solutions 28
1.5.1 Structural disorder and diffusion of oxide ions
in CexZr1–xO2 solid solutions 30
1.5.2 Visualization of diffusion pathway of oxide ions
in the ceria–yttria solid solution Ce1−xYxO2−x/2 33
1.5.3 Structural disorder and diffusion of oxide ions
in Ce0.8R0.2O1.9 solid solutions 35
1.6 Conclusions and Discussion 38
Acknowledgments 40
References 40

Chapter 2 Understanding Ceria-Based Catalytic Materials: An


Overview of Recent Progress 47
Juan José Delgado, Eloy del Río, Xiaowei Chen,
Ginesa Blanco, José María Pintado, Serafín Bernal
and José Juan Calvino
2.1 Introduction 47
2.2 A Model for the Nanostructure, Pre-Treatment
Conditions and Redox Behaviour in High-Temperature
Aged Ceria–Zirconia Catalysts 51
2.2.1 Oxygen storage capacity of CeO2−ZrO2 vs CeO2 51
2.2.2 Effect on OSC of high-temperature redox cycles 53
2.2.3 Chemical and structural investigation of OSC
changes after high-temperature redox cycling 55
2.2.4 CO-OSC of ceria–zirconia bare oxides 73

b1469_FM.indd x 4/8/2013 12:43:14 PM


b1469 Catalysis by Ceria and Related Materials

Contents xi

2.3 Chemical Characterization of Cerium-Based


Oxide-Supported Gold Catalysts 78
2.3.1 Hydrogen adsorption on cerium-based
oxide-supported gold catalysts 82
2.3.2 CO adsorption on cerium-based oxide-supported
gold catalysts 99
Acknowledgments 116
References 117

Chapter 3 Investigation of the Oxygen Storage and Release


Kinetics of Model and Commercial
Three-Way Catalytic Materials
by Transient Techniques 139
Angelos M. Efstathiou and Stavroula Y. Christou
3.1 Scope 139
3.2 Introduction 140
3.3 Theoretical Background of Transient Techniques
Used in Oxygen Storage and Release Kinetic Studies 142
3.3.1 Pulse-injection technique 143
3.3.2 Dynamic oxygen storage capacity (DOSC) 152
3.3.3 Modeling of OSC 157
3.3.4 Oxygen isotopic exchange 158
3.3.5 Is OSC in three-way catalysts an equilibrium-
controlled process? 168
3.4 Case Studies 169
3.4.1 Deterioration of oxygen storage and release kinetics
in CeO2 and CexZr1−xO2−δ due to chemical poisoning 169
3.4.2 Deterioration of the oxygen storage and release
kinetics in CeO2 and CexZr1−xO2−δ induced
by thermal ageing 186
3.4.3 Regeneration of the oxygen storage and release
properties of aged commercial and model
three-way catalysts 202
Acknowledgments 212
References 212

b1469_FM.indd xi 4/8/2013 12:43:14 PM


b1469 Catalysis by Ceria and Related Materials

xii Contents

Chapter 4 Interaction of Nitrogen Oxides with Ceria-Based


Materials 223
Avelina García-García and Agustin Bueno-López
4.1 Introduction 223
4.2 N2O Interaction with Ceria-Based Materials: N2O
Decomposition on Rh/Ceria Catalysts 226
4.3 NOx Interaction with Ceria-Based Materials:
Ceria-Catalyzed Oxidation of NO to NO2 233
4.4 Conclusions 243
References 244

Chapter 5 Atomistic Modelling of Ceria Nanostructures:


Introducing Structural Complexity 247
Dean C. Sayle and Thi X. T. Sayle
5.1 Methodology 247
5.1.1 Potential models 248
5.1.2 Theoretical methods 248
5.2 Introducing Microstructure 249
5.2.1 The pristine parent material 249
5.2.2 Point defects: oxygen vacancies 250
5.2.3 Oxygen transport 250
5.2.4 Surfaces 252
5.2.5 Surface oxygen vacancies 253
5.2.6 Surface oxygen transport 254
5.2.7 Dislocations 256
5.2.8 Grain boundaries 257
5.2.9 Supported thin films and interfaces 259
5.3 Simulating ‘The Dirt’ 260
5.3.1 Ceria nanotubes 261
5.3.2 Outlook 263
5.4 Nanocatalysis 264
5.5 Simulating Synthesis 265
5.5.1 Simulated annealing and Monte Carlo 265
5.5.2 Atom deposition 266
5.5.3 Simulated crystallisation 269

b1469_FM.indd xii 4/8/2013 12:43:14 PM


b1469 Catalysis by Ceria and Related Materials

Contents xiii

5.5.4 Ceria nanoparticles and nanorods 273


5.5.5 Mesoporous ceria 281
5.6 Ab Initio Methods 287
5.7 Outlook 289
Acknowledgements 289
References 290

Chapter 6 Two-Dimensional and Three-Dimensional


Ceria-Based Nanoarchitectures 295
Zhen-Xing Li, Wei Feng, Chao Zhang,
Ling-Dong Sun, Ya-Wen Zhang,
and Chun-Hua Yan
6.1 Introduction 295
6.2 Synthesis and Assembly of Ceria-Related Nanomaterials 297
6.2.1 Synthesis of ceria-based nanomaterials 297
6.2.2 Assembly of ceria-based nanomaterials 311
6.3 Properties and Applications 322
6.3.1 Catalysis 323
6.3.2 Sensors 332
6.3.3 Biomedical applications 336
6.3.4 Optical applications 342
6.4 Conclusions and Outlook 345
References 347

Chapter 7 Core-Shell-Type Materials Based on Ceria 361


Matteo Cargnello, Raymond J. Gorte,
and Paolo Fornasiero
7.1 Introduction 361
7.2 Synthesis of Core-Shell Materials Based on Ceria 364
7.2.1 Co-precipitation 365
7.2.2 Microemulsion-based synthesis 369
7.2.3 Other methods 373
7.3 Applications 376
7.3.1 Water gas shift reaction (WGSR) 376
7.3.2 Gas-phase oxidation reactions 384
7.3.3 Selective organic transformations 387

b1469_FM.indd xiii 4/8/2013 12:43:14 PM


b1469 Catalysis by Ceria and Related Materials

xiv Contents

7.3.4 Fuel-cell catalysts 389


7.4 Conclusions 392
Acknowledgements 392
References 392

Chapter 8 New Developments in Ceria-Based Mixed


Oxide Synthesis and Reactivity in Combustion
and Oxidation Reactions 397
Benjaram M. Reddy, Thallada Vinod Kumar
and Naga Durgasri
8.1 Introduction 397
8.2 Synthetic Approaches 399
8.2.1 Co-precipitation 400
8.2.2 Hydrothermal method 404
8.2.3 Solvothermal method 407
8.2.4 Sol-gel method 409
8.2.5 Microemulsions 414
8.2.6 Microwave method 416
8.2.7 Sonochemical method 420
8.2.8 Combustion synthesis 422
8.2.9 Impregnation method 424
8.2.10 Electrochemical method 426
8.2.11 Ball milling 428
8.2.12 Spray pyrolysis 430
8.3 Catalytic Combustion and Oxidation Applications 431
8.3.1 Catalytic combustion of VOCs 432
8.3.2 Soot combustion and oxidation 438
8.3.3 CO oxidation 444
8.3.4 Hydrocarbon oxidation 453
8.4 Conclusions 454
Acknowledgments 455
References 455

b1469_FM.indd xiv 4/8/2013 12:43:14 PM


b1469 Catalysis by Ceria and Related Materials

Contents xv

Chapter 9 Design and Modeling of Active Sites


in Metal–Ceria Catalysts for the Water Gas Shift
Reaction and Related Chemical Processes 465
Jose A. Rodriguez
9.1 Introduction: Ceria-Based Catalysts
for the Water Gas Shift Reaction 465
9.2 Active Phase of Au–CeO2, Cu–CeO2, and Pt–CeO2
Powder Catalysts 467
9.3 Water Gas Shift Reaction on Au and Cu Nanoparticles
Supported on a Well-defined CeO2(111) Surface 470
9.4 Inverse CeOx/Au(111) and CeOx/Cu(111) Catalysts
and the Chemistry Associated with the Water Gas
Shift Reaction 474
9.5 Water Gas Shift Reaction on M/CeOx/TiO2(110)
Surfaces (M=Cu, Au, or Pt) 485
9.6 Conclusion 491
Acknowledgements 492
References 492

Chapter 10 Ceria-Based Gold Catalysts: Synthesis,


Properties, and Catalytic Performance
for the WGS and PROX Processes 497
Donka Andreeva, Tatyana Tabakova
and Lyuba Ilieva
10.1 Introduction 497
10.2 Preparation of Gold–Ceria Catalysts 499
10.2.1 Preparation of ceria and doped ceria
as supports for gold catalysts 499
10.2.2 Deposition of gold 506
10.3 Characterization of Gold–Ceria Catalysts 511
10.3.1 Dispersion and oxidation state of supported gold 511
10.3.2 State of ceria and modified ceria supports 515
10.4 WGS Catalytic Activity on Gold–Ceria Catalysts 531
10.4.1 WGS reaction on gold–ceria 531
10.4.2 WGS reaction over gold based on doped ceria 533

b1469_FM.indd xv 4/8/2013 12:43:14 PM


b1469 Catalysis by Ceria and Related Materials

xvi Contents

10.5 Active Sites and Mechanism of the WGS Reaction


Over Ceria-Based Gold Catalysts 539
10.6 PROX Activity Over Gold–Ceria Catalysts 549
10.7 Concluding Remarks 556
Acknowledgments 557
References 557

Chapter 11 Ceria-Based Formulations for Catalysts


for Diesel Soot Combustion 565
Eleonora Aneggi, Carla de Leitenburg
and Alessandro Trovarelli
11.1 Introduction 565
11.1.1 Background 565
11.1.2 Removal of soot in diesel engine emissions 568
11.1.3 Experimental approaches 569
11.2 Overview of the Use of Ceria for Soot Oxidation 572
11.3 Bare Ceria Formulations 575
11.4 Ceria-Based Solid Solutions 582
11.5 Formulations Based on Metal-Doped Ceria 590
11.6 Reactions in a NOx/O2 Atmosphere 600
11.7 Conclusions 610
Acknowledgments 611
References 611

Chapter 12 Ceria and its Use in Solid Oxide Cells


and Oxygen Membranes 623
Christodoulos Chatzichristodoulou,
Peter T. Blennow, Martin Søgaard,
Peter V. Hendriksen, and Mogens B. Mogensen
12.1 Introduction 623
12.2 The Chemistry and Physics of Ceria in Brief 625
12.2.1 Defects in pure and doped ceria 626
12.2.2 Lattice parameters of pure, doped
and reduced ceria 627

b1469_FM.indd xvi 4/8/2013 12:43:14 PM


b1469 Catalysis by Ceria and Related Materials

Contents xvii

12.3 Defect Chemistry and Thermodynamic Properties


of Pure and Doped Ceria 631
12.3.1 Defects in pure and doped ceria 631
12.3.2 Defect thermodynamic properties of pure
and doped ceria 636
12.4 Ionic and Electronic Conductivity of Pure
and Doped Ceria 652
12.4.1 Pure ceria 652
12.4.2 Effect of doping on the ionic and electronic
conductivity of ceria 658
12.5 Use of Ceria as an Electrolyte in Solid Oxide
Fuel Cells 675
12.5.1 Solid oxide fuel cells 675
12.5.2 Theoretical analysis 681
12.5.3 Overview of performance 691
12.5.4 Technological programs 699
12.5.5 Interdiffusion barrier layer 700
12.6 Use of Ceria for Oxygen Membranes 701
12.6.1 Oxygen 701
12.6.2 Gas separation membranes 702
12.6.3 Applications of oxygen gas separation
membranes 705
12.6.4 Ceria membranes 707
12.6.5 Mechanical problems of ceria in membranes 725
12.6.6 Concluding remarks on ceria membrane
performance 726
12.7 Ceria in Solid Oxide Fuel Cell Electrodes 728
12.7.1 Introduction 728
12.7.2 Ceria-based electrodes 736
12.7.3 Ceria nanoparticles in SOFC electrodes 741
12.7.4 Mechanistic aspects of ceria in SOC electrodes 748
12.7.5 Concluding remarks 754
Acknowledgments 755
References 755

b1469_FM.indd xvii 4/8/2013 12:43:14 PM


b1469 Catalysis by Ceria and Related Materials

xviii Contents

Chapter 13 Transformation of Oxygenated Compounds


Derived from Biomass into Valuable
Chemicals Using Ceria-Based Solid Catalysts 783
Laurence Vivier and Daniel Duprez
13.1 Introduction 783
13.2 Dehydrogenation of Ethanol to Acetaldehyde 783
13.3 Synthesis of Carbonates 786
13.4 Reaction of Glycerol 789
13.4.1 Oxidation 790
13.4.2 Dehydration 791
13.4.3 Hydrogenolysis 793
13.4.4 Transesterification 794
13.5 Selective Hydrogenation of C=O Bonds
of α,β-Unsaturated Aldehydes 795
13.6 C–C Coupling Reactions 799
13.6.1 Ketonization 799
13.6.2 Aldol condensation 802
13.7 Oxidation of Sugar Derivatives 805
13.8 Conclusions 806
References 807

Chapter 14 Ceria-Based Catalysts for Air Pollution


Abatement 813
Anna Maria Venezia, Leonarda Francesca Liotta,
Giuseppe Pantaleo, and Alessandro Longo
14.1 Introduction 813
14.2 Oxidation of CO 814
14.2.1 Nanostructured CeO2 815
14.2.2 Ceria-based oxide catalysts 818
14.2.3 Ceria-supported noble-metal catalysts 826
14.3 Oxidation of VOCs 830
14.3.1 Ceria and ceria-based mixed oxides 831
14.3.2 Ceria-supported metal catalysts 842
14.4 Methane Combustion 850
14.4.1 Ceria-based oxide catalysts 851
14.4.2 Noble-metal catalysts 854

b1469_FM.indd xviii 4/8/2013 12:43:14 PM


b1469 Catalysis by Ceria and Related Materials

Contents xix

14.5 Reduction of NO with CO 861


14.5.1 Ceria-promoted metal-supported catalysts 863
14.5.2 Ceria-supported metal and oxide catalysts 865
14.6 Conclusion 871
Acknowledgements 872
References 873

Index 881

b1469_FM.indd xix 4/8/2013 12:43:14 PM


b1469 Catalysis by Ceria and Related Materials

This page intentionally left blank

b1469_FM.indd xx 4/8/2013 12:43:14 PM


b1469 Catalysis by Ceria and Related Materials

CHAPTER 1

CRYSTAL AND ELECTRONIC


STRUCTURES, STRUCTURAL DISORDER,
PHASE TRANSFORMATION, AND PHASE
DIAGRAM OF CERIA–ZIRCONIA AND
CERIA-BASED MATERIALS
Masatomo Yashima
Department of Chemistry and Materials Science, Graduate School of
Science and Engineering, Tokyo Institute of Technology, 2-12-1-W4-17,
O-okayama, Meguro-ku, Tokyo, 152-8551, Japan

1.1 Phase Diagram of the Ce–O System and Structural


Properties of Fluorite-Type CeO2
Cerium exhibits both the +4 and +3 oxidation states.1 The change of
oxidation state between +4 and +3 is the key for the catalysis of ceria-
based materials. Thus, the phase diagram of the Ce–O system is
important for ceria-based catalysts. The Ce–O system contains a
number of crystalline phases (Fig. 1.1).1–3 Table 1.1 shows some of
the phases and their space groups in the Ce–O system.3–10 There is a
disagreement in the literature concerning both the composition
and the structure of intermediate phases.
Cerium dioxide CeO2 has a fluorite-type structure (space group
Fm3m),3–7 which is named after the mineral form of calcium fluo-
ride. Figures 1.2(a)–(c) show the crystal structure of cerium dioxide
CeO2. The cerium and oxygen atoms are located at the 4a 0,0,0 and
8c 1/4,1/4,1/4 sites, respectively.

b1469_Ch-01.indd 1 4/8/2013 12:25:03 PM


b1469 Catalysis by Ceria and Related Materials

2 M. Yashima

Figure 1.1 Phase diagrams of the Ce–O system. (a) Phase diagram of CeOy, 1.7 ≤
y ≤ 2.0, temperature range 630 ≤ T ≤ 1270 K.1, 2 (b)–(d) Calculated phase diagrams of
the Ce–O system. Mole fraction of O is x in Ce1-xOx (b) 0.0 ≤ x ≤ 1.0, 500 ≤ T ≤ 4500 K,
(c) 0.55 ≤ x ≤ 0.70, 500 ≤ T ≤ 3000 K, (d) 0.63 ≤ x ≤ 0.67, 600 ≤ T ≤ 1200 K.3 Reprinted
with permission from Elsevier.

The coordination number of the oxygen atoms is four (CN = 4)


and the structure consists of OCe4 tetrahedra (Fig. 1.2(b)). The
coordination number of the Ce cations is eight (CN = 8). The crystal
structure consists of CeO8 cubes (Fig. 1.2(c)).
The unit-cell parameter a of CeO2 increases with temperature
(Fig. 1.3(a)).5 The atomic displacement parameters of both Ce and

b1469_Ch-01.indd 2 4/8/2013 12:25:03 PM


b1469 Catalysis by Ceria and Related Materials

Structure, Disorder, Transformation, and Phase Diagram of Ceria-Based Materials 3

Table 1.1 Some phases and space groups of cerium oxides.

Phase (symbol) in Fig. 1.1 O/Ce ratio n m Space group References


CeO2-x (F, F’, α, α’) 2−x ∞ 1 Fm3m 3–7
Ce11O20 (δ) 1.818 11 1 P1 3, 4
Ce62O112 (p1) 1.806–1.808 62 6 — 3, 8
Ce40O72 (ε) 1.800 40 4 — 3, 9
Ce19O34 (p2, M19) 1.790 19 2 triclinic 3, 8
Ce9O16 (ζ) 1.778 9 1 — 3, 9
Ce7O12 (τ, ι) 1.714 7 1 R3 3, 4
Ce3O5±x (C) (5 ± x)/3 11 1 R3 3, 4
Ce2O3-x (A) (3 − x)/2 4 1 P 3m1 3, 10

Note: The numbers n and m refer to the stoichiometry compound CenO2n–2m.

O (B(Ce) and B(O)) also increase with temperature (Fig. 1.3(b)). 5


The higher B(O) values (B(O) > B(Ce)) suggest that the oxygen
anions are mobile compared with the Ce cations.
The large thermal vibration of oxygen atoms in CeO2 is confirmed
also by the nuclear-density distribution (Fig. 1.3(c)).6,7 The nuclear-
density map indicates that the distribution of oxide ions increases with
an increase of temperature.7 The bulges towards the <111> direction
are clearly observed for the oxide ions in CeO2 indicating anisotropic
thermal motion.6,7 This complicated disorder of the oxide ions is
responsible for ionic conduction in ceria-based electrolytes. Two pos-
sible diffusion paths, in the direction <111> and along the c axis, have
been suggested for CeO2 over a wide temperature range. These fea-
tures have also been observed in the fast oxide-ion conductor Bi2O3.12
Both CeO2 and Bi2O3 have a cubic fluorite-type structure, suggesting
that this disorder and the possible diffusion paths are characteristic of
compounds with a cubic fluorite-type structure.
Figure 1.1 shows that the fluorite-structured CeO2-x has a solid
solution phase. The unit-cell parameter of CeO2-x increases with an
increase of defect concentration x1,13 because the ionic radius of
Ce3+ is higher than that of Ce4+. For a coordination number of 8
(CN = 8), the ionic radius of Ce3+ after Shannon14 is r(Ce3+,8) =
1.143 Å, while that of Ce4+ is smaller, r(Ce4+,8) = 0.97 Å.

b1469_Ch-01.indd 3 4/8/2013 12:25:04 PM


b1469 Catalysis by Ceria and Related Materials

4 M. Yashima

Figure 1.2 (a) Crystal structure of the cubic fluorite-type ceria CeO2. (b) Crystal
structure of ceria with OCe4 tetrahedra. (c) Crystal structure of ceria with CeO8
cubes. These figures were drawn with the program VESTA.11

1.2 Phases and Structure of Zirconia and Ceria–Zirconia


Solid Solutions
Ceria–zirconia materials exhibit high catalytic activity and are used
for exhaust gas cleaning.15 To understand the catalytic activity of
ceria–zirconia materials, a knowledge of the phases and the crystal
structure of zirconia and ceria–zirconia solid solutions is
important.

b1469_Ch-01.indd 4 4/8/2013 12:25:04 PM


b1469 Catalysis by Ceria and Related Materials

Structure, Disorder, Transformation, and Phase Diagram of Ceria-Based Materials 5

Figure 1.3 Temperature dependence of (a) unit-cell parameter a and (b) atomic
displacement parameters of the cubic fluorite-type CeO2.5 Reprinted with permis-
sion from Yashima et al.5 Copyright 2003 Elsevier. (c) Nuclear-density distribution of
the fluorite-type CeO2 at 1770 K.6, 7 Reprinted with permission from Yashima et al.6
Copyright 2004 American Institute of Physics.

1.2.1 Phase transition and crystal structure of zirconia


Zirconia ceramics are used for a variety of applications as catalysts,
structural materials, and electrolytes for solid-oxide fuel cells. Zirconia
(zirconium dioxide ZrO2) exhibits a phase transition sequence,16–19
1400 K 2642 K 2983 K
Monoclinic (m)  Tetragonal (t) ↔ Cubic (c) ↔ Liquid

b1469_Ch-01.indd 5 4/8/2013 12:25:04 PM


b1469 Catalysis by Ceria and Related Materials

6 M. Yashima

The monoclinic-tetragonal (m-t) phase transition is first order and


has a martensitic nature. There is a large hysteresis between the transi-
tion temperatures on heating and on cooling.17 The t-to-m phase transi-
tion is induced by stresses around cracks, which improves the toughness
of zirconia ceramic materials. The tetragonal-cubic phase transition
does not exhibit large hysteresis. The nature of the tetragonal-cubic
phase transition has not been established yet, probably due to the dif-
ficulty of experiments at high temperatures. The cubic-tetragonal
phase transition is important for the formation of the tetragonal phase.
Figures 1.4(a) and 1.4(b) show parts of the baddeleyite-type
structure of zirconium dioxide ZrO2. This structure belongs to the
monoclinic space group P 21/c and has a distorted fluorite-type struc-
ture.19 The coordination number (CN) of the Zr cations is seven,
which is smaller than that of the fluorite-type oxide (CN = 8). This
smaller value (CN = 7) is attributable to the smaller size of Zr cations
(r(Zr4+,8) = 0.84 Å) compared to the criterion deduced from
Pauling’s first rule (r(A,8) ≥1.079 Å) for A cations with a CN of eight.
The high-temperature tetragonal phase also has a distorted
fluorite-type structure with the P42/mcm space group (Figs. 1.4(c)

Figure 1.4 (a) Crystal structure of the monoclinic ZrO2 projected on to the (100)
plane. The ZrO7 polyhedra are drawn from crystallographic data.19 Schematic cat-
ion coordination polyhedra of (b) monoclinic, (c) tetragonal, and (d) cubic phases.
Panel (b–d) reprinted with permission from Yashima et al.19 Copyright 1995
American Physical Society.

b1469_Ch-01.indd 6 4/8/2013 12:25:05 PM


b1469 Catalysis by Ceria and Related Materials

Structure, Disorder, Transformation, and Phase Diagram of Ceria-Based Materials 7

Figure 1.5 Crystal structure of tetragonal zirconia and relation between tetragonal
(t) cell and pseudo-cubic fluorite (F) cell. The closed and shaded open circles
denote the cations and anions, respectively. An arrow in an open circle indicates the
displacement of oxygen atoms along the c axis. The thick blue dashed lines indicate
a primitive tetragonal cell and thin solid lines stand for two pseudo-fluorite cells.20,21

and 1.5). The cubic-to-tetragonal phase transition is accompanied by


oxygen displacement from the regular 8c position in the Fm3m sym-
metry (arrows in Figs. 1.4(c) and 1.5) and by elongation of the c axis.
The coordination number of a Zr cation in tetragonal ZrO2 is eight
as well as in cubic fluorite-type ZrO2 (Figs. 1.4(c) and 1.4(d)).

1.2.2 Crystal structure of monoclinic CexZr1–xO2 and tetragonal-


monoclinic phase transformation in the CeO2–ZrO2 system
In this section the phases and their crystal structures in composition-
ally homogeneous CeO2–ZrO2 samples are described (Table 1.2).
Undoped pure ZrO2 has a monoclinic P 21/c baddeleyite-type
structure, which is a distorted fluorite-type structure. The monoclinic
CexZr1–xO2 solid solution forms in the compositional region of 0 ≤ x ≤
0.12 (Table 1.2 and Fig. 1.6). The phase boundary is strongly
dependent on the sample processing and grain size because these
factors influence the nucleation, growth, and kinetics of the

b1469_Ch-01.indd 7 4/8/2013 12:25:05 PM


b1469 Catalysis by Ceria and Related Materials

8 M. Yashima

Table 1.2 Phases in bulk compositionally homogeneous samples of the CeO2–


ZrO2 system at room temperature.

Compositional
range of Phases and Space Unit-cell
CexZr1-xO2 crystal system group References parameters
0 ≤ x < 0.08a Monoclinic (m) P 21/c 17, 19 a ≠ b ≠ c,
β > 90ο
0.08a ≤ x ≤ 0.12a m+t P 21/c + 17, 19
P 42/nmc
0.12a < x ≤ 0.20 Tetragonal (t) P 42/nmc 17, 19 c/aF > 1b
0.20 < x ≤ 0.60a Tetragonal (t′) P 42/nmc 22–24 c/aF > 1
0.60a ≤ x < 0.65a t′ + t″ P 42/nmc + 22–24
P 42/nmc
0.65a ≤ x < 0.90 Tetragonal (t″) P 42/nmc 22–24 c/aF = 1
0.90 ≤ x Cubic (c) F m3m 22–24 c/aF = 1

Notes: a The compositional range is dependent on the processing and grain size. baF is the unit-
cell parameter a of the pseudo-cubic fluorite cell.

Figure 1.6 Compositional dependence of (a) unit-cell parameters of monoclinic


(am, bm, cm) and tetragonal (at, ct) CexZr1-xO2 and (b) β angle of monoclinic (βm) and
tetragonal (βt = 90°) CexZr1-xO2 at room temperature. See Yashima et al.19 for the
meaning of the symbols. Reprinted with permission from Yashima et al.19 Copyright
1995 American Physical Society.

b1469_Ch-01.indd 8 4/8/2013 12:25:06 PM


b1469 Catalysis by Ceria and Related Materials

Structure, Disorder, Transformation, and Phase Diagram of Ceria-Based Materials 9

tetragonal-monoclinic phase transformation.16 In the monoclinic


phase, the unit-cell parameters am and cm increase considerably while
bm shows a rather small increase with increasing Ce content x
(Fig. 1.6(a)). The angle βm decreases with increasing x (Fig. 1.6(b)),
which indicates a decrease of the monoclinic distortion. The unit-cell
parameters and the atomic positions discontinuously change during
the monoclinic-tetragonal phase change when the Ce content x is
changed. The discrete change is consistent with the first-order mono-
clinic-tetragonal phase transition. The thermal ellipsoids show a large
anisotropy, and they correlate with the direction of movement of
oxide ions during the monoclinic-tetragonal phase transformation.19
The crystal structure of the monoclinic phase approaches that of the
tetragonal phase with an increase of CeO2 content x: (1) the am value
approaches the bm one with increasing x (Fig. 1.6(a)), (2) the angle βm
decreases with x (Fig. 1.6(b)), and (3) all positional parameters of the
monoclinic CexZr1-xO2 approach those of the tetragonal structure
(Fig. 1.7). Figure 1.8 shows the compositional dependence of the
unit-cell volume of CexZr1-xO2, which indicates a volume change Vm/
(2Vt) = 1.05. The volume change is responsible for the tetragonal-to-
monoclinic transformation toughening of zirconia ceramic materials.

Figure 1.7 Compositional dependence of positional parameters x(O1) and y(O1)


of monoclinic CexZr1-xO2 at room temperature.19 Reprinted with permission from
Yashima et al.19 Copyright 1995 American Physical Society.

b1469_Ch-01.indd 9 4/8/2013 12:25:07 PM


b1469 Catalysis by Ceria and Related Materials

10 M. Yashima

Figure 1.8 Compositional dependence of unit-cell volume of monoclinic (Vm) and


tetragonal (2Vt) CexZr1-xO2 at room temperature.19 Reprinted with permission from
Yashima et al.19 Copyright 1995 American Physical Society.

Figure 1.9 Compositional dependence of (a) the transition temperature and


(b) the transition enthalpy (kJ/(mol of cation)) between the monoclinic and
tetragonal phases in CexZr1-xO2.17

Yashima et al. used differential scanning calorimeter (DSC) meas-


urements to study the tetragonal-monoclinic phase transformation of
compositionally homogeneous CexZr1-xO2 materials (0 ≤ x ≤ 0.12).17
The tetragonal-monoclinic transition temperature, enthalpy, and
entropy decrease with an increase of ceria concentration x (Fig. 1.9).
The transition enthalpy ∆H t-m is a linear function of x in CexZr1-xO2.17

b1469_Ch-01.indd 10 4/8/2013 12:25:07 PM


b1469 Catalysis by Ceria and Related Materials

Structure, Disorder, Transformation, and Phase Diagram of Ceria-Based Materials 11

∆H t-m = 5.64 – 36.3 x (kJ/(mol of cation))

The decrease of ∆H t-m with ceria concentration (Fig. 1.9(b)) is con-


sistent with that of βm (Fig. 1.6(b)) and of bm–am (Fig. 1.6(a)). The
decrease of ∆H t-m is attributable to the approach of the crystal
structure of the monoclinic phase to that of the tetragonal phase
(Figs 1.6 and 1.7).

1.2.3 Phase identification and crystal structure of tetragonal phase in the


compositionally homogeneous CexZr1–xO2 − t, t ′, and t ″ forms
Phases and unit-cell parameters of bulk CexZr1-xO2 materials have been
studied by numerous researchers (Fig. 1.10).17, 19, 22–39 Table 1.2 shows
the phases in bulk compositionally homogeneous CexZr1-xO2 samples.
There are three forms of the tetragonal phase in CexZr1-xO2 materials:
t is the stable tetragonal form at a high temperature, while t′ and t″ are
metastable or non-equilibrium forms. The axial ratio c/aF of the t and
t′ forms is higher than unity (c/aF > 1) while the c/aF ratio of t″ form
is equal to unity (c/aF = 1); the subscript F denotes the pseudo-fluorite

Figure 1.10 Compositional dependence of unit-cell parameters of tetragonal and


cubic phases in CexZr1-xO2. Data are taken from Yashima et al. (1993),22, 25 Yoshimura
et al. (1990),40 and Tsukuma et al. (1984).26 aF, cF and VF are the unit-cell parameters
a and c and unit-cell volume for the pseudo-fluorite lattice.

b1469_Ch-01.indd 11 4/8/2013 12:25:08 PM


b1469 Catalysis by Ceria and Related Materials

12 M. Yashima

lattice. The t″ form is tetragonal because the oxygen atoms are dis-
placed along the c axis from the regular 8c position of the cubic phase
(Figs. 1.4(c) and 1.5). The three forms, t, t′, and t″, belong to the same
tetragonal phase with P42/nmc symmetry. The existence of the t″ form
is evidenced by X-ray23, 24 and neutron24 powder diffraction, Raman
scattering,23 transmission electron microscopy (TEM),35 and density
functional theory (DFT) calculations.41 The t′–t″ phase boundary is
detectable by high-angular-resolution synchrotron X-ray diffractome-
try (Fig. 1.11).24 In bulk samples of CexZr1-xO2 for x = 0.65 and x = 0.70,
no peak splitting between the 004F and 400F reflections was detected,

Figure 1.11 High-angular-resolution synchrotron X-ray powder diffraction pro-


files of compositionally homogeneous CexZr1-xO2.24 Reprinted with permission from
Yashima et al.24 Copyright 1998 American Institute of Physics.

b1469_Ch-01.indd 12 4/8/2013 12:25:08 PM


b1469 Catalysis by Ceria and Related Materials

Structure, Disorder, Transformation, and Phase Diagram of Ceria-Based Materials 13

Figure 1.12 Neutron diffraction profiles around the 112F reflection of CexZr1-xO2.24
Reprinted with permission from Yashima et al.24 Copyright 1998 American Institute
of Physics.

while the 112F reflection and tetragonal Raman bands were clearly
observed (Figs. 1.1224 and 1.1323). The subscript F denotes the pseudo-
fluorite lattice. These results indicate the existence of the t″ form. The
t″ form appears not only in bulk CexZr1-xO223,24 but also in nano-crystal-
line samples.36 In the nano-sized CexZr1-xO2, it is difficult to distinguish
t″ from t′ due to peak broadening. For example, peak splitting
between the 004F and 400F reflections is not detectable in nano-sized
Ce0.2Zr0.8O242 and Ce0.4Zr0.6O2.36
Figure 1.10 shows the composition x dependence of the unit-cell
parameters of the compositionally homogeneous CexZr1-xO2 materi-
als.22,25 The unit-cell parameters and volume of tetragonal and cubic
CexZr1-xO2 increase with an increase of ceria concentration x. The
increase of unit-cell volume with x is attributable to the larger ionic

b1469_Ch-01.indd 13 4/8/2013 12:25:08 PM


b1469 Catalysis by Ceria and Related Materials

14 M. Yashima

Figure 1.13 Raman spectra of compositionally homogeneous ceria–zirconia solid


solutions CexZr1-xO2 where x is from 0.5 to 0.9.23 The tetragonal phase has six active
Raman modes of A1g + 3Eg + 2B1g. Numbers in the figure are the six Raman bands.
The asterisk (*) denotes the band of structural disorder. The third band (arrow) is
a good indication of the tetragonal phase. Reprinted with permission from Yashima
et al.23 Copyright 1994 John Wiley & Sons.

radius of Ce4+ (r(Ce4+,8) = 0.97 Å)14 compared to that of Zr4+


(r(Zr4+,8) = 0.84 Å).14 The aF and cF parameters approach each other
with an increase of x, which leads to a decrease of the axial ratio c/
aF. Yashima41 studied the crystal structures of tetragonal and cubic
CexZr1-xO2 materials using DFT calculations, which also gave a
decrease of the axial ratio c/aF with ceria composition x.
Figures 1.14(a) and 1.14(b) show the compositional depend-
ence of the axial ratio c/aF and oxygen displacement along the c axis
from the cubic fluorite regular position, respectively.24 Both the c/aF
ratio and oxygen displacement decrease with ceria content x. The
c/aF ratio discretely becomes unity (c/aF = 1) at x = 0.6, which sug-
gests that the t′–t″ transformation is a first-order transition. Thus,
the t′–t″ boundary is dependent on processing and grain size.
Oxygen displacement (Fig. 1.14(b)) and the tetragonal
Raman band (the third band of Fig. 1.13) are zero around x = 0.9 for
CexZr1-xO2, which indicates the t″–cubic transformation.23,24 The com-
position of the t″–cubic phase boundary (x = 0.9) in bulk CexZr1-xO223,24
agrees with that of nano-sized CexZr1-xO2 as reported by Lamas et al.36

b1469_Ch-01.indd 14 4/8/2013 12:25:09 PM


b1469 Catalysis by Ceria and Related Materials

Structure, Disorder, Transformation, and Phase Diagram of Ceria-Based Materials 15

Figure 1.14 Variation of (a) axial ratio c/aF and (b) oxygen displacement along
the c axis from the regular 8c position in CexZr1-xO2.24 Reprinted with permission
from Yashima et al.24 Copyright 1998 American Institute of Physics.

1.3 Phase Diagram and Partitionless Transformation


in the CeO2–ZrO2 System
1.3.1 Experimental and calculated equilibrium phase diagram
of the CeO2–ZrO2 system
The phase diagrams of ZrO2-containing systems have been investi-
gated by numerous researchers; however, there are many discrepan-
cies and misunderstandings in the literature.16 There are various
factors, which make the phase diagram and structural change com-
plicated:16 (1) a variety of metastable states due to sluggish kinetics
(such as a slow diffusion of cations, nucleation and growth processes,
a strong dependency of the kinetics on the starting material, process-
ing, etc.), (2) a variety of metastable states due to partitionless
(sometimes called as “diffusionless”) phase transformations, (3) the
effect of other species on metastability and chemical reactions, and

b1469_Ch-01.indd 15 4/8/2013 12:25:09 PM


b1469 Catalysis by Ceria and Related Materials

16 M. Yashima

Figure 1.15 Experimental equilibrium phase diagram of the CeO2–ZrO2 system.50


liq.: liquid, c: cubic, t: tetragonal, m: monoclinic. See Yashima et al.50 for an explana-
tion of the symbols. Reprinted with permission from Yashima et al.50 Copyright 1994
John Wiley & Sons.

(4) the difficulty of conducting experiments at high temperatures


above 1900 K and below 1200 K.
The equilibrium phase diagram of the CeO2–ZrO2 system has also
been investigated by many researchers.43–51 Figure 1.15 shows the
experimental equilibrium phase diagram in the CeO2–ZrO2 system
proposed by our group.50 We studied the low-temperature phase equi-
libria ranging from 1273 K to 1473 K by annealing CexZr1-xO2 samples
in a Na2B2O7–NaF flux. We confirmed the phase separation into
stable monoclinic (m) and tetragonal (t) phases and the two-phase
region (m + t). A eutectoid reaction from t to (m + c) was found to
occur at 1328 K where the equilibrium compositions of the t, m, and
c phases were x = 0.112, 0.009, and 0.84 in CexZr1-xO2, respectively.
The solubility limit of the monoclinic phase is small (x = 0.009).

b1469_Ch-01.indd 16 4/8/2013 12:25:10 PM


b1469 Catalysis by Ceria and Related Materials

Structure, Disorder, Transformation, and Phase Diagram of Ceria-Based Materials 17

Figure 1.16 Calculated51 (solid line) and measured52 (A),53 (B),54 and (C)55 mean
heat capacity of CeO2. Mean heat capacity is defined as (H(T) − H(294 K))/
(T − 294), where H is the enthalpy in J (mol of cation)−1 K−1. Reprinted with permis-
sion from Du et al.51 Copyright 1994 Elsevier.

Du et al.51 studied the equilibrium phase diagram in the CeO2–


ZrO2 system using the CALPHAD (CALculation of PHAse Diagram)
method. First the lattice stability of CeO2 was evaluated to reproduce
the measured data for the mean heat capacity52–55 (Fig. 1.16).51 For
example, the lattice stability of cubic CeO2, G c-CeO2 (298 < T < 3083 K)
was determined to be
G c-CeO2 = –1116114 + 429.345T − 72.0653T lnT
− 0.0040536T 2 + 583870T −1.
Figure 1.17 shows the calculated phase diagram of the CeO2–ZrO2
system, which is consistent with our experimental diagram (Fig. 1.15).

1.3.2 Temperature-induced structural change of tetragonal CexZr1–xO2


and partitionless tetragonal–cubic phase transformation in
compositionally homogeneous CexZr1–xO2 materials
The partitionless tetragonal-cubic phase transformation temperature
in bulk CexZr1-xO2 has been studied by ex situ X-ray powder diffraction
experiments (Figs. 1.18 and 1.1956). Compositionally homogenous t′

b1469_Ch-01.indd 17 4/8/2013 12:25:10 PM


b1469 Catalysis by Ceria and Related Materials

18 M. Yashima

Figure 1.17 Calculated equilibrium phase diagram of the CeO2–ZrO2 system using
the CALPHAD method.51 liq.: liquid, c: cubic, t: tetragonal, m: monoclinic. Symbols
stand for experimental data. See Du et al.51 for an explanation of the symbols.
Reprinted with permission from Du et al.51 Copyright 1994 Elsevier.

CexZr1-xO2 samples were annealed at a high temperature (Tanneal) for


a short time (typically 3 min.), quenched to room temperature (RT)
and then investigated by X-ray powder diffractometry at RT. As shown
in Fig. 1.19(a), peak splitting between the 004F and 400F reflections
of compositionally homogeneous Ce0.65Zr0.35O2 decreases with an
increase of annealing temperature Tanneal and disappears at Tanneal =
1304 K. The c/aF ratio does not change with annealing temperature
Tanneal at lower temperatures, while near the tetragonal–cubic phase
transition temperature, c/aF decreases with Tanneal and becomes unity
(Fig. 1.19(b)), which indicates the finishing temperature of the t′–(t″
or cubic) phase transformation. The t′–(t″ or cubic) transformation
temperature decreases with an increase of the CeO2 concentration x
in CexZr1-xO2 (Figs. 1.19(b) and the alternating long and single short
dash line in Fig. 1.18). The t″–cubic phase transition has been

b1469_Ch-01.indd 18 4/8/2013 12:25:10 PM


b1469 Catalysis by Ceria and Related Materials

Structure, Disorder, Transformation, and Phase Diagram of Ceria-Based Materials 19

Figure 1.18 Phase diagram of compositionally homogeneous CexZr1-xO2 where the


partitionless transformation temperatures of CexZr1-xO2 are plotted.17,23,24,56 This is not
an equilibrium but a non-equilibrium phase diagram. This diagram has incorrectly
been described as a metastable phase diagram.16 liq.: liquid, c: cubic, t: equilibrium
tetragonal form, t′: metastable/non-equilibrium tetragonal form with axial ratio c/aF
> 1, t″: metastable/non-equilibrium tetragonal form with axial ratio c/aF = 1.

investigated by high-temperature Raman scattering.23 The intensity


of the third Raman band characteristic of the tetragonal phase in
t″-Ce0.8Zr0.2O2 decreases with increasing temperature and disappears
between 673 and 773 K. The t″–cubic phase transition of Ce0.8Zr0.2O2
occurs reversibly during heating and cooling. The cubic–t″ transition
is induced by oxygen displacement from the regular fluorite 8c
1/4,1/4,1/4 position along the c axis. The t″–cubic transition tem-
perature is indicated by the alternating long and two short dashes
line in Fig. 1.18. At some temperature, these t″–t′ and t″–cubic phase
boundaries in the non-equilibrium phase diagram (Fig. 1.18) are
located near the equilibrium phase boundary between the (t+c)

b1469_Ch-01.indd 19 4/8/2013 12:25:11 PM


b1469 Catalysis by Ceria and Related Materials

20 M. Yashima

Figure 1.19 (a) Annealing temperature dependence of ex situ X-ray powder diffrac-
tion profiles of compositionally homogeneous Ce0.65Zr0.35O2 obtained by quenching
after annealing the tetragonal t′ phase for a short time (typically 3 min.).56 (b) Variation
of the axial ratio c/aF of compositionally homogeneous CexZr1-xO2 with annealing tem-
perature.25,56,57 In Ce0.65Zr0.35O2, the t′ (c/aF > 1) and t″ or cubic (c/aF = 1) phases coexist
below the finishing point of the t′-to-(t″ or cubic) transformation. Reprinted with
permission from Yashima et al.56 Copyright 1993 John Wiley & Sons.

two-phase and cubic single-phase regions in the equilibrium phase


diagram (Fig. 1.15). A thermodynamic model of a simple regular
solution where the interaction parameter is independent of tempera-
ture and ceria content was not able to reproduce the location of the
t′–t″ and t″–cubic phase boundaries.57 In contrast, we successfully
reproduced the location using Landau’s phenomenology where the
tetragonal–cubic phase transition is approximated as second order.
The structural change of bulk t′-Ce0.5Zr0.5O2 and the t′–cubic
transformation were investigated by in situ neutron powder diffrac-
tion measurements in air up to 1829 K (Fig. 1.20).21 The isotropic
atomic displacement parameters of Ce and Zr atoms U(Ce,Zr) and
oxygen atoms U(O) increase with temperature, with U(O) being
higher than U(Ce,Zr) suggesting higher diffusivity of oxide ions. The
c/aF axial ratio for the tetragonal Ce0.5Zr0.5O2 increases in the range
from 296 K to 1036 K and decreases in the range from 1293 K to
1543 K, and becomes unity between 1543 K and 1829 K (Fig. 1.20(a)).
The displacement of oxygen atoms along the c axis in the tetragonal

b1469_Ch-01.indd 20 4/8/2013 12:25:11 PM


b1469 Catalysis by Ceria and Related Materials

Structure, Disorder, Transformation, and Phase Diagram of Ceria-Based Materials 21

Figure 1.20 Temperature dependence of (a) axial ratio c/aF and (b) oxygen dis-
placement from the fluorite regular position along the c axis in the compositionally
homogeneous Ce0.5Zr0.5O2.21 Open and closed circles are for the t′ and cubic phases,
respectively. Reprinted with permission from Wakita et al.21 Copyright 2007 IUCr.

Ce0.5Zr0.5O2 increases in the range from 296 K to 1036 K and decreases


in the range from 1293 K to 1543 K, and becomes 0.0 Å between
1543 K and 1829 K (Fig. 1.20(b)). These results indicate that the
cubic–t′ phase transition between 1543 K and 1829 K is accompanied
by oxygen displacement along the c axis and an increase of the c/aF
axial ratio from unity. The transition temperature obtained by an
in situ neutron diffraction study21 agrees with that from the ex situ
X-ray powder diffraction measurements.56 Near the tetragonal–cubic
transition temperature, further detailed investigation of the precise
crystal structure of CexZr1-xO2 is needed to determine the mechanism
of the cubic–tetragonal phase transformation. However, the CexZr1-xO2
material decomposes into more stable (t + c) two phases during
measurements longer than 5 min where the t and c have lower and
higher ceria concentrations, respectively.
The temperature dependence of nano-sized CexZr1-xO2 was stud-
ied by in situ synchrotron X-ray powder diffractometry.39 Due to
peak broadening for the nano-sized CexZr1-xO2 and weak X-ray scat-
tering by oxygen, the reported tetragonal–cubic transformation
temperature might be incorrect.

b1469_Ch-01.indd 21 4/8/2013 12:25:11 PM


b1469 Catalysis by Ceria and Related Materials

22 M. Yashima

Figure 1.21 (a) Schematic phase diagram, showing the equilibrium phase bounda-
ries (solid lines) and partitionless transformation temperatures (dotted line and
dashed lines). (b)–(e) Gibbs free energy-composition (G-x) diagram of the CeO2–
ZrO2 system.16 Schematic G-x diagrams at temperatures (b) T = T 1, (c) T = T 2, and
(d) T = T 3. (e) Schematic G-x diagram for constant temperature to explain the
formation of the t′ form by annealing the precursor containing Ce and Zr cations.42
(a)–(d) Reprinted with permission from Yashima et al.16 Copyright 1996 Elsevier.
(e) Reprinted with permission from Yashima et al.42 Copyright 1994 John Wiley & Sons.

1.3.3 Thermodynamics of the formation of t′-CexZr1–xO2


and nano-sized t′-CexZr1–xO2
It is useful to use a schematic Gibbs free energy-composition (G-x)
diagram to discuss the mechanism of the formation of a metastable
phase.16 Figures 1.21(b)–(d) show schematic G-x diagrams for the
CeO2–ZrO2 system, which can be used to understand the formation
of non-equilibrium m′, t′, and t″ forms. The high-temperature cubic
phase (W0 in Fig. 1.21(a)) is cooled to a temperature T 1 (Y1 in
Fig. 1.21(a)) and it can be seen that the Gibbs free energy of the
cubic phase (W1 in Fig. 1.21(b)) is higher than that of the t′ form

b1469_Ch-01.indd 22 4/8/2013 12:25:12 PM


b1469 Catalysis by Ceria and Related Materials

Structure, Disorder, Transformation, and Phase Diagram of Ceria-Based Materials 23

with the same chemical composition (Y1 in Fig. 1.21(b)). Thus, a


partitionless cubic–t′ phase transformation occurs (W1 to Y1 in
Fig. 1.21(b)). The t′ form (Y1 in Fig. 1.21(b)) is unstable compared
to the stable t+c two-phase state (E1 + H1 in Fig. 1.21(b)).
A schematic G-x diagram (Fig. 1.21(e)) is useful also for discuss-
ing the formation mechanism of nano-sized t′-CexZr1-xO2 obtained
by heating the precursors derived from chemical processing.42 The
amorphous precursor has a high G value (P in Fig. 1.21(e)) and
transforms into the more stable t′ phase during heating at lower
temperatures or for a shorter time where the phase separation into
t+c two phases does not occur (Y in Fig. 1.21(e)). Figure 1.22 shows

Figure 1.22 Raman spectra of (a) the precursor, and (b) compositionally homo-
geneous nano-sized and (c) bulk ceria–zirconia solid solutions Ce0.2Zr0.8O2 obtained
by heating the precursor at 723 K (b) and at 1863 K (c).42 The tetragonal phase has
six Raman active modes of A1g + 3Eg + 2B1g. Numbers in the figure are the six
Raman bands. The asterisk (*) denotes the band of structural disorder. The third
band is a good indication of the tetragonal phase. Reprinted with permission from
Yashima et al.42 Copyright 1994 John Wiley & Sons.

b1469_Ch-01.indd 23 4/8/2013 12:25:12 PM


b1469 Catalysis by Ceria and Related Materials

24 M. Yashima

the Raman spectrum of Ce0.2Zr0.8O2 synthesized by the polymerized


complex method (Pechini method).42
The precursor (Fig. 1.22(a)) and samples heated at 723 K
(Fig. 1.22(b)) and at 1863 K (Fig. 1.22(c)) show tetragonal Raman
bands, although X-ray diffraction profile of the sample heated at
723 K does not exhibit peak splitting between the 004F and 400F
reflections due to peak broadening of the nano-sized crystallites.
Várez et al. obtained similar conclusions through a detailed investi-
gation using X-ray powder diffraction, energy-dispersive spectrome-
try (EDS), TEM, nuclear magnetic resonance (NMR), and Raman
scattering.58 The TEM-EDS technique detected small compositional
inhomogeneities in their samples. Using X-ray diffraction, the coex-
istence of the t′, t, and c phases was clearly observed in the sample
heated at 1173 K due to the phase separation from t′ to t+c.58 This
phase decomposition is explained by a lowering of the Gibbs free
energy (the Y to (E+H) = F change in Fig. 1.21(e)42).

1.4 Electronic Structure and Chemical Bonding of


Zirconia, Ceria, and Ceria–Zirconia Solid Solutions
DFT calculations are useful for investigating the ground state.59 A
number of researchers have utilized the DFT technique to study the
electronic structures of CeO2, Cex Zr1-xO2 solid solutions, and
ZrO2.41,60–71 Assuming the invalid cubic structure, Rodriguez et al.60
examined the compositional dependence of the crystal structure of
Cex Zr1-xO2 by the DFT method. The tetragonal Cex Zr1-xO2 phase is
much more important for automobile exhaust catalysts.41

1.4.1 Optimized structure and valence-electron density of tetragonal


zirconia and ceria
Figures 1.23(a) and (b) show the optimized crystal structure and
isosurface of the corresponding valence-electron density distribu-
tion of tetragonal zirconia (Zr8O16 supercell).41 The optimized unit-
cell parameters41 agree with experimental values.72 The average

b1469_Ch-01.indd 24 4/8/2013 12:25:12 PM


b1469 Catalysis by Ceria and Related Materials

Structure, Disorder, Transformation, and Phase Diagram of Ceria-Based Materials 25

Figure 1.23 (a) Optimized structure and (b) isosurface of valence-electron den-
sity at 0.0133 Å−3 with valence-electron density distributions on the (100), (010),
and (001) planes of a supercell (2 × 2 × 2) of tetragonal zirconia ZrO2.41 (c)
Optimized structure and (d) isosurface of valence-electron density at 0.0133 Å−3
with valence-electron density distributions on the (100), (010), and (001) planes of
a (2 × 2 × 1) supercell of tetragonal ceria CeO2.41 The optimized structure of CeO2
is of the cubic fluorite type. Yellow and green spheres are Ce and Zr atoms, respec-
tively. Pink/purple spheres in (a) and (c) are oxygen atoms. The arrows along the
c axis in (a) indicate the direction of oxygen displacement from the ideal fluorite
8c position. On the color scale for the valence-electron density, 0 and 100% corre-
spond to 0.0133 and 0.148 Å−3, respectively. Reprinted with permission from
Yashima.41 Copyright 2009 ACS.

value of the calculated z coordinate of the oxygen atoms41 also


agrees with the experimental z value.72 The oxygen atoms are dis-
placed from the regular position (z = 1/4) (Fig. 1.23(a)).41 The
oxygen displacement yields longer and shorter Zr–O interatomic
distances. Covalent bonds between the Zr and O atoms are observed
for the shorter Zr–O bond in the calculated valence-electron density
map (Fig. 1.23(b)). These features from the DFT calculations are
consistent with the crystal structure and electron density of the

b1469_Ch-01.indd 25 4/8/2013 12:25:13 PM


b1469 Catalysis by Ceria and Related Materials

26 M. Yashima

Figure 1.24 (a) Refined crystal structure and (b) isosurface of experimental elec-
tron density of tetragonal zirconium oxide.73 The arrows in (a) show oxygen dis-
placement along the c axis from the regular fluorite 8c position. Reprinted with
permission from Yashima and Tsunekawa.73 Copyright 2006 IUCr.

zirconium oxide obtained through synchrotron X-ray powder dif-


fraction experiments (Figs. 1.24(a,b)).73
Figures 1.23(c) and (d) show the optimized crystal structure and
isosurface of the corresponding valence-electron density distribution
of ceria (Ce8O16 supercell).41 The optimized unit-cell parameters
agree with the experimental values. The relation c = 2 a and opti-
mized atomic coordinates indicate that the ceria has a cubic fluorite-
type structure. The calculated valence-electron density distribution
of ceria is consistent with the experimental data.74 In comparison
with the Zr–O bonds in zirconia (Fig. 1.23(b)), the Ce–O bond in
ceria is more ionic (Fig. 1.23(d)).

1.4.2 Optimized structure and valence-electron density of tetragonal


ceria–zirconia solid solutions
Figure 1.25 shows the optimized crystal structure and isosurface of
the corresponding valence-electron density distribution of the tetrag-
onal (t′) ceria–zirconia solid solution Ce0.5Zr0.5O2.41 The calculated
tetragonal unit-cell parameters41 agree with experimental values.21

b1469_Ch-01.indd 26 4/8/2013 12:25:14 PM


b1469 Catalysis by Ceria and Related Materials

Structure, Disorder, Transformation, and Phase Diagram of Ceria-Based Materials 27

Figure 1.25 (a) Optimized crystal structure and (b) yellow isosurface of valence-
electron density at 0.0133 Å−3 with valence-electron density distributions on the
(100), (010), and (001) planes of the (2 × 2 × 1) supercell of tetragonal (t′) ceria-
zirconia solid solution Ce0.5Zr0.5O2.41 The arrows in (a) and (b) show oxygen
displacement along the c axis from the regular fluorite 8c position. Reprinted with
permission from Yashima.41 Copyright 2009 ACS.

The average value of the calculated z coordinate of an oxygen atom


(0.214)41 also agrees with the experimental z value (0.2190).21
The oxygen atoms are displaced from the regular position (z = 1/4)
(Fig. 1.25(a)). Oxygen displacement yields longer and shorter cat-
ion–oxygen interatomic distances. Covalent bonds between Zr and O
atoms in Ce0.5Zr0.5O2 are observed for the shorter Zr–O bond in the
calculated valence-electron density map (Fig. 1.25(b)) as well as in
ZrO2 (Fig. 1.23(b)). The covalent bond is formed by the overlap of
the Zr 3d and O 2p orbitals as shown in the partial density of states
(Fig. 1.26). These features are also obtained for other compositions
CexZr1-xO2 (x = 1/8, 1/4, 3/8, 5/8, and 3/4).41
The theoretical calculations41 demonstrated that the axial ratio
and oxygen displacement from the regular position of the opti-
mized structures of CexZr1-xO2 decrease with increasing CeO2 con-
tent (closed circles and solid lines in Fig. 1.27), which agrees well
with those determined by diffraction experiments of ceria–zirconia
solid solutions prepared using a high-temperature process24 (open
squares in Fig. 1.27). The tetragonal t″ form with an axial ratio of

b1469_Ch-01.indd 27 4/8/2013 12:25:15 PM


b1469 Catalysis by Ceria and Related Materials

28 M. Yashima

Figure 1.26 Partial density of states of a Zr and an O atom in a (2 × 2 × 1) supercell


of tetragonal (t′) ceria-zirconia solid solution Ce0.5Zr0.5O2.41 Reprinted with permis-
sion from Yashima.41 Copyright 2009 ACS.

unity is strongly thought to be stable in Ce0.875Zr0.125O2, in compari-


son with the tetragonal t′ form with an axial ratio greater than unity
and the cubic phase. As described in this section, recent develop-
ments in DFT calculations have enabled theoretical studies of the
crystal and electronic structures of ceria–zirconia CexZr1-xO2 materi-
als. However, further developments and more accuracy in the
calculations from first principles are required for studying the oxy-
gen-deficient CexZr1-xO2-δ, larger supercells, long-range ordering,
modulated structures, phase diagrams, the phase-transition mecha-
nism, and the temperature dependence of the crystal structure.

1.5 Structural Disorder and Diffusion Pathways of Oxide


Ions in Ceria–Zirconia and Ceria–Rare-Earth Oxide
Solid Solutions
Solid materials with high oxide-ion conductivity have attracted consid-
erable attention owing to their many applications such as in solid-
oxide fuel cells (SOFCs), batteries, catalysts, gas sensors, and oxygen
separation membranes. An atomic-scale understanding of the oxide-
ion diffusion mechanism in oxide-ion conductors has provided one of
the major challenges in the field of condensed matter physics, solid-
state chemistry, and materials science. A combined technique using
the maximum-entropy method (MEM), MEM-based pattern fitting,
the Rietveld method, and high-temperature neutron/synchrotron

b1469_Ch-01.indd 28 4/8/2013 12:25:17 PM


b1469 Catalysis by Ceria and Related Materials

Structure, Disorder, Transformation, and Phase Diagram of Ceria-Based Materials 29

Figure 1.27 (a) Axial ratio c/aF −1 and (b) average value of oxygen displacement
from the ideal fluorite 8c position of ceria–zirconia solid solutions CexZr1-xO2 as a
function of CeO2 content x in (2 × 2 × 1) supercells. Solid circles and open squares
denote theoretical41 and experimental24 data, respectively.

X-ray powder diffractometry has enabled the visualization of the struc-


tural disorder and diffusional pathway of mobile ions in various ionic
conductors.75–80 Classical Rietveld analysis is able to refine the atomic
displacement parameters, which describe the static and dynamic

b1469_Ch-01.indd 29 4/8/2013 12:25:17 PM


b1469 Catalysis by Ceria and Related Materials

30 M. Yashima

positional disorder of mobile oxide ions.81 This section describes the


application of these techniques to the disorder and diffusional path-
way of oxide ions in ceria–zirconia and ceria-based ionic conductors.

1.5.1 Structural disorder and diffusion of oxide ions


in CexZr1–xO2 solid solutions
Diffusion of oxide ions is an important step in the oxygen storage/
release process of ceria-based catalysts.82 We investigated the crystal
structure and structural disorder of compositionally homogeneous
CexZr1-xO2 and CeO2 by synchrotron and neutron powder diffrac-
tometry, the Rietveld method, and MEM.6,7,32,74,83 The isotropic
atomic displacement parameter of oxygen atoms U(O) in tetragonal
CexZr1-xO2 and cubic CeO2 is higher than that of Ce and Zr atoms
U(Ce,Zr) in the whole compositional range 0.12< x <1.0, U(O) >
U(Ce,Zr) (Fig. 1.28), which suggests the higher diffusivity of oxide

Figure 1.28 Compositional dependence of the isotropic atomic displacement


parameters of Ce and Zr atoms U(Ce,Zr) and O atoms U(O) in tetragonal
CexZr1–xO2 solid solutions and cubic CeO2 (x = 0.12 to 1) at 299 K.32 Open circles
and triangles denote U(Ce,Zr) and U(O) of the tetragonal CexZr1-xO2 solid solu-
tions, respectively. Closed circle and triangle are U(Ce) and U(O) for cubic CeO2,
respectively. Reprinted with permission from Yashima et al.32 Copyright 2009
American Institute of Physics.

b1469_Ch-01.indd 30 4/8/2013 12:25:17 PM


b1469 Catalysis by Ceria and Related Materials

Structure, Disorder, Transformation, and Phase Diagram of Ceria-Based Materials 31

Figure 1.29 Relation between the isotropic atomic displacement parameter of O


atoms U(O) and the diffusion coefficient of O atoms in CexZr1-xO2 solid solutions.32
The solid line was obtained by a least-squares fit (quadratic polynomial). Reprinted
with permission from Yashima et al.32 Copyright 2009 American Institute of Physics.

ions compared to cations. U(O) for CexZr1-xO2 increases with an


increase of CeO2 content x in the compositional range 0.12 ≤ x ≤ 0.5,
while U(O) decreases with x in 0.5 ≤ x ≤ 1.0 (Fig. 1.28). Thus, the
Ce0.5Zr0.5O2 composition has the highest U(O) value in CexZr1-xO2
solid solutions (x = 0.12, 0.4, 0.5, 0.6, 0.7, 0.8, and 1.0), suggesting
higher bulk diffusivity of the oxide ions in Ce0.5Zr0.5O2 compared
with those of other compositions.
Figure 1.29 shows the relation between the isotropic atomic dis-
placement parameter of oxygen atoms U(O) and the oxygen diffu-
sion coefficient D(O).32 The D(O) data were taken from the
literature.84 D(O) increases with an increase of U(O).32 Thus, the
atomic displacement parameter of oxide ions U(O) is a good indica-
tor of oxygen diffusivity. High U(O) values indicate high dynamic
thermal vibration or high static positional disorder, which enhance
the movement of oxide ions across the unit cell. High U(O) is
regarded as the structural origin of high oxide-ion diffusivity.32,83
Figure 1.30 shows the electron-density distributions of tetrago-
nal Cex Zr1-xO2 (x = 0.12 and 0.5) and cubic CeO2.32 The electron-
density distribution of CeO232 is consistent with the literature.74 The
oxide ions in Ce0.5Zr0.5O2 are spread over a wide area compared to
CexZr1-xO2 (x = 0.12 and 1.0), which suggests higher bulk diffusivity

b1469_Ch-01.indd 31 4/8/2013 12:25:18 PM


b1469 Catalysis by Ceria and Related Materials

32 M. Yashima

Figure 1.30 Electron density distributions on the (100) planes of (a) Ce0.12Zr0.88O2
and (b) Ce0.5Zr0.5O2 with black contours in the range from 4.0 to 20.0 Å−3 (steps of
2.0 Å−3) at 299 K, 1/2 ≤ z ≤ 1. (c) Electron density distribution on the (110) plane
of cubic CeO2 with black contours in the range from 4.0 to 20.0 Å−3 (steps of 2.0 Å−3)
at 299 K, 1/2 ≤ z ≤ 1.32 Reprinted with permission from Yashima et al.32 Copyright
2009 American Institute of Physics.

of the oxide ions of Ce0.5Zr0.5O2. The atomic displacement parame-


ter of oxygen atoms in Ce0.5Zr0.5O2, U(O) = 0.0276(9) Å2, is higher
than in CeO2, U(O) = 0.0092(6) Å2, (Fig. 1.28),32 which is consistent
with the results at 1832 K.83 Boaro et al.84 reported that the oxide-ion
diffusion coefficient of Ce0.5Zr0.5O2 is higher than that of CeO2,
which corresponds to the higher U(O) in Ce0.5Zr0.5O2. Similarly the
atomic displacement parameter of the oxygen atoms U(O) of
CexZr1-xO2 increases with CeO2 content x in the compositional range
0.12 ≤ x ≤ 0.5, and the U(O) of CexZr1-xO2 decreases with x in 0.5 ≤ x
≤ 1.0 at 299 K (Fig. 1.28).32 Thus, the U(O) of Ce0.5Zr0.5O2 as well as
the oxide-ion diffusion coefficient has a maximum value in 0.12 < x <
1.0. The electron-density distributions of CexZr1-xO2 (x = 0.12, 0.5,
and 1) in Fig. 1.30 indicate that the spatial distribution of oxide ions
in Ce0.5Zr0.5O2 is greater than those of the others. Therefore, oxygen
diffusivity is higher when the U(O) is higher and the spatial distribu-
tion of the oxide ions is greater.
We studied the nuclear density distributions of Ce0.5Zr0.5O2 at
1832 K and of CeO2 at 1826 K by in situ neutron diffraction and
MEM (Fig. 1.31).83 The nuclear density distribution of oxide ions in
Ce0.5Zr0.5O2 (Fig. 1.31(a)) indicates the large positional disorder of
oxide ions, spreading over a wide area and a shift to the <111> direc-
tions. Possible diffusion paths of the oxide ions can be seen along the

b1469_Ch-01.indd 32 4/8/2013 12:25:18 PM


b1469 Catalysis by Ceria and Related Materials

Structure, Disorder, Transformation, and Phase Diagram of Ceria-Based Materials 33

Figure 1.31 Nuclear density distributions on the (110) planes of (a) cubic
Ce0.5Zr0.5O2 at 1832 K and (b) CeO2 at 1826 K with black contours in the range from
0.7 to 5.0 fm Å−3 (steps of 0.5 fm Å−3).83 The dotted lines (A) and (B) with arrows
are possible diffusion paths in the <100> and <110> directions, respectively.
Reprinted with permission from Wakita and Yashima.83 Copyright 2008 American
Institute of Physics.

<100> and <110> directions ((A) and (B) in Fig. 1.31, respectively).
The spatial distribution of oxide ions in Ce0.5Zr0.5O2 (Fig. 1.31(a)) is
greater than that of CeO2 (Fig. 1.31(b)). The greater degree of posi-
tional disorder in Ce0.5Zr0.5O2 is responsible for the higher diffusivity
of the oxide ions leading to higher catalytic activity.

1.5.2 Visualization of diffusion pathway of oxide ions in the ceria–yttria


solid solution Ce1–xYxO2–x/2
The development of ceria-based materials requires an understand-
ing of the diffusion mechanism at an atomic scale. For this purpose
it is important to study the diffusional pathway of mobile oxide ions
in a crystal lattice at high temperatures where the materials work
efficiently and the pathway can be visualized clearly. The crystal
structures of CeO2, CeO2-δ, and Ce1-xYxO2-x/2 have been investigated
at high temperature by various researchers.5–7,74,85–89 We studied the
crystal structure, disorder, and diffusion path of an oxide-ion con-
ductor, the yttria-doped ceria solid solution Ce0.93Y0.07O1.96.89 We
chose the composition Ce0.93Y0.07O1.96 because it has the maximum

b1469_Ch-01.indd 33 4/8/2013 12:25:19 PM


b1469 Catalysis by Ceria and Related Materials

34 M. Yashima

Figure 1.32 Isosurface of nuclear density at 0.003 fm Å−3 (0.15 < x < 0.3) with nuclear-
density distribution on the bc plane at x = 0.3 of Ce0.93Y0.07O1.96 obtained by the combina-
tion of Rietveld refinement and the maximum-entropy method of neutron powder
diffraction data measured at 296 K for (c), and at 1707 K for (b) and (d).89

electrical conductivity in ceria–yttria solid solutions.90 The ceria–


yttria solid solution Ce0.93Y0.07O1.96 has a cubic defect fluorite-type
structure as shown by the refined structure in Fig. 1.32(a).89 The
MEM map illustrated the positional disorder of the oxide ions (Fig.
1.32(b)–(d)), compared with the structural model obtained by
Rietveld analysis (Fig. 1.32(a)). The conventional simple model of
atomic spheres is not appropriate for describing the positional dis-
order of oxide ions. The spatial distribution of oxide ions at 1707 K
is larger than at 296 K, corresponding to the higher atomic displace-
ment parameters.89 There are two types of bulge in the MEM
nuclear-density distribution map. One bulge expands along the
<100> direction. The other exists along the <110> direction. These
directions for the oxide ions are on opposite side of the Ce and Y

b1469_Ch-01.indd 34 4/8/2013 12:25:20 PM


b1469 Catalysis by Ceria and Related Materials

Structure, Disorder, Transformation, and Phase Diagram of Ceria-Based Materials 35

cations. The bulges are attributable to repulsion between the cations


and the oxide ions. Similar anisotropic features in the nuclear den-
sity distribution have been observed in ceria CeO2.6,7 The shift of
oxide ions in the <111> direction has been reported also in bismuth
oxide compounds12,75,76,91–95 and the β′ fluorite phase in the Mg-Zr-
O-N system.96 These materials have the same defect cubic fluorite-
type structure as ceria-based materials. Therefore, this feature is
common to cubic fluorite-type materials.75,76,89

1.5.3 Structural disorder and diffusion of oxide ions in


Ce0.8R0.2O1.9 solid solutions
In order to investigate the correlation between the atomic displace-
ment parameter of oxygen atoms and oxygen diffusivity (Fig. 1.33),
the crystal structures of rare-earth oxide-doped ceria materials,
Ce0.8R0.2O1.9 (where R is one of the rare earth metals La, Nd, Sm, Gd,
Y, or Yb) were studied by Rietveld analyses of high-resolution synchro-
tron X-ray powder diffraction data taken in air in situ at 303 K, 681 K,
and 948 K (d > 0.38 Å).81 As a reference material, non-doped ceria
CeO2 was also examined by the same method. These materials were
confirmed to have a defect fluorite-type structure (Fig. 1.33(a)). The
unit-cell parameters a of Ce0.8R0.2O1.9 a(R) and CeO2 a(CeO2)
increase with increasing temperature. There is a relation between the
unit-cell parameters a(R) of Ce0.8R0.2O1.9 at constant temperature:
a(La) > a(Nd) > a(Sm) > a(Gd) > a(CeO2) > a(Y) > a(Yb). The a(R)
parameter increases with an increase of the ionic radius r(R) at each
temperature of 303 K, 681 K, and 948 K. r(R) for a coordination
number of eight after Shannon14 was used. The isotropic atomic dis-
placement parameters of cations UC(R) and anions UO(R) increase
with temperature, which indicates an increase of thermal vibration
and dynamic positional disorder (Fig. 3.33(c)), compared with those
at the lower temperature of 303 K (Fig. 3.33(b)). At a constant tem-
perature, the atomic displacement parameters of a cation in
Ce0.8R0.2O1.9 UC(R) are higher than in CeO2 UC(CeO2), which indi-
cates that the static and dynamic positional disorder of cations in
Ce0.8R0.2O1.9 is also larger compared with CeO2. Similarly, at a

b1469_Ch-01.indd 35 4/8/2013 12:25:21 PM


b1469 Catalysis by Ceria and Related Materials

36 M. Yashima

Figure 1.33 (a) Crystal structure of the defect fluorite-type ceria–rare-earth oxide
solid solution Ce0.8R0.2O1.9.81 (b), (c), (d) Schematic diagrams of potential curves for
oxide ions, with vibrating oxide ions.81 Figures (b) and (c) show small and large
thermal vibrations of oxide ions, respectively. Large thermal vibrations can lead to
oxide ions hopping to a neighboring oxygen site. Figure (d) schematically indicates
that large static disorder has higher and lower energy barriers. The higher activation
energy in (d) means lower mobility of oxide ions, which leads to lower oxide-ion
conductivity. Reprinted with permission from Yashima et al.81 Copyright 2010 ACS.

constant temperature, the atomic displacement parameters of anions


of Ce0.8R0.2O1.9 UO(R) are higher than CeO2 UO(CeO2), which indi-
cates that the static and dynamic positional disorder of oxide ions in
Ce0.8R0.2O1.9 is also larger than in CeO2. The greater dynamic disorder
of oxide ions in Ce0.8R0.2O1.9 is the structural origin of the higher
ionic conductivities of Ce0.8R0.2O1.9 than for undoped CeO2.
Figure 1.34(a) shows the dopant ionic radius r(R) dependence
of the atomic displacement parameter UO(R) of anions in Ce0.8R0.2O1.9.
UO(R) decreases with increasing ionic radius r(R) from R = Yb to R =
Gd, while UO(R) increases with r(R) from R = Gd to R = La. Thus,
UO(R) has a minimum at R = Gd in the range from R = Yb to R = La.

b1469_Ch-01.indd 36 4/8/2013 12:25:21 PM


b1469 Catalysis by Ceria and Related Materials

Structure, Disorder, Transformation, and Phase Diagram of Ceria-Based Materials 37

Figure 1.34 (a) Variation of atomic displacement parameter UO(R) for an anion
with the ionic radius of a dopant cation r(R) for Ce0.8R0.2O1.9 (R = La, Nd, Sm, Gd,
Y, and Yb) measured at 948 K.81 (b) Variation of ionic conductivity σ with UO(R) for
Ce0.8R0.2O1.981 measured at 1073 K.97 Reprinted with permission from Yashima et al.81
Copyright 2010 ACS.

The mismatch between dopant rare-earth oxide and host ceria gives
rise to the static positional disorder, which leads to the higher
atomic displacement parameters. The static positional disorder
reflects the local distortion or lattice strain due to the mismatch
between host CeO2 and dopant RO1.5, which prevents fast oxide-ion
movement across the crystal lattice (Fig. 1.33(d)). Therefore, oxide-
ion conductivity decreases with increasing UO(R) as shown in Fig.
1.34(b). On the other hand, we would expect that an increase of
dynamic disorder and effective index would improve oxide-ion con-
ductivity.81 UO(R) increases with an increase of effective index from
R = Gd to R = La,81 which indicates the possibility that the bulk
oxide-ion conductivity of Ce0.8R0.2O1.9 increases with an increase of
dopant ionic radius r(R) from R = Gd to R = La.81 Further careful
structural investigation of ceria-based materials is needed to deter-
mine both the dynamic and static components of the atomic dis-
placement parameter of oxygen atoms.

b1469_Ch-01.indd 37 4/8/2013 12:25:22 PM


b1469 Catalysis by Ceria and Related Materials

38 M. Yashima

1.6 Conclusions and Discussion


In this chapter, I reviewed the crystal structure, electronic structure,
structural disorder, phase transformation, and phase diagrams of
ceria–zirconia and ceria-based materials.
Phases, crystal structure, and phase diagrams in the CeO2–ZrO2
system have been studied by many researchers. But, there are many
contradictions in the literature. Equilibrium phases are formed by
diffusional transformation. In partitionless and compositionally
homogeneous samples, the cation-diffusionless phase transforma-
tions are key to understanding the phases of the CeO2–ZrO2 system.
I described not only the conventional equilibrium phase diagram
(Figs. 1.15 and 1.17) but also the non-equilibrium phase diagram of
the CeO2–ZrO2 system, with partitionless phase transformation
temperatures (Fig. 1.18). This non-equilibrium phase diagram has
been called a metastable phase diagram.16 The Gibbs free energy-
composition (G -x) diagram is useful for understanding the forma-
tion of m′, t′, and t″ forms and partitionless phase transformations.
In situ neutron-diffraction studies have indicated that the partition-
less tetragonal t′–cubic phase transformation is accompanied by an
elongation of the length of the c axis compared to the a axis and by
oxygen displacement along the c axis (Fig. 1.20).21 As discussed in
Section 1.3.2, further detailed investigation of the precise crystal
structure of Cex Zr1–xO2 near the tetragonal–cubic transition tem-
perature is needed to understand the mechanism of the partition-
less cubic–tetragonal phase transformation. However, Cex Zr1–xO2
decomposes into the more stable t form with low ceria content and
the cubic phase with high ceria concentration during measurements
for times longer than 5 min. To distinguish the cation-diffusionless
phase transition from the diffusional phase separation, a tempera-
ture-time-transformation diagram is useful (Fig. 1.35).22,25
Recent precise structural analyses of ceria–zirconia and ceria–
rare-earth-oxides have demonstrated the positional disorder and
diffusional pathway of oxide ions at high temperatures. In a ceria–
yttria material, the oxide ions diffuse along the <100> directions
(Fig. 1.32). We found that the diffusion coefficient increases with

b1469_Ch-01.indd 38 4/8/2013 12:25:23 PM


b1469 Catalysis by Ceria and Related Materials

Structure, Disorder, Transformation, and Phase Diagram of Ceria-Based Materials 39

Figure 1.35 Temperature-time-transformation (TTT) diagram containing both


(c or t″)-to-t′ partitionless phase transition and (c or t″ or t′)-to-(c + t) diffusional
transformation in Ce0.5Zr0.5O2.22,25 Reprinted with permission from Yashima et al.22
Copyright 1993 John Wiley & Sons.

increasing atomic displacement parameter for oxygen atoms U(O)


in ceria–zirconia materials (Fig. 1.29). The spatial distribution of
oxide ions in Ce0.5Zr0.5O2 is greater than that of CeO2 (Figs. 1.30 and
1.31), which is consistent with the higher bulk diffusivity of oxide
ions in Ce0.5Zr0.5O2. The greater atomic displacement parameter and
large spatial distribution of oxide ions in Ce0.5Zr0.5O2 are possible
factors for its higher catalytic activity.
The observed atomic displacement parameters include not only
the dynamic component but also the static one. Static disorder can
lead to lower oxygen diffusivity, while dynamic disorder increases
ionic conductivity (Fig. 1.33). The ionic conductivity and diffusivity
of oxide ions in Ce0.8R0.2O1.9 decreases with an increase of the atomic
displacement parameter (Fig. 1.34). As discussed in Section 1.5.3,
further structural studies of the atomic displacement parameters in

b1469_Ch-01.indd 39 4/8/2013 12:25:23 PM


b1469 Catalysis by Ceria and Related Materials

40 M. Yashima

ceria-based materials are required to distinguish the dynamic com-


ponent from static disorder and to establish the structure-diffusivity
correlation.
Although many researchers have studied the phase change and
crystal structure of oxygen deficient Cex Zr1−xO2−δ,98–105 these have not
yet been established. The structure change during the oxygen stor-
age/release process is also important for future study. Structure
studies of nano-sized ceria-based materials will be important in the
future. For this purpose, the characterization of local structures is
required.

Acknowledgments
I would like to express special thanks to all my co-authors and col-
laborators mentioned in the literature. Some of this work was finan-
cially supported by the Ministry of Education, Culture, Sports,
Science and Technology of Japan, through Grant-in-Aid for Scientific
Research (A) and (B), and Challenging Exploratory Research
(Nos. 24246107, 21360318, 23655190). This work was carried out
under the Joint-use Research Program for Neutron Scattering,
Institute for Solid State Physics (ISSP), the University of Tokyo, at
the Research Reactor JRR-3, JAEA (Proposal Nos. 8766, 7768, 6723,
6721, 5681, 5682, 4688, 4689, 3653). The synchrotron data were col-
lected at the BL-4B2 beam line of the Photon Factory under the
projects 2008G084, 2006G264, 2005G158 and at the BL02B2 beam
line of SPring-8 (Project Nos. 2011B1995 and 2011A1442).

References
1. Trovarelli, A., in Catalysis by Ceria and Related Materials, ed. A. Trovarelli,
Imperial College Press, London, (2002), pp. 15–50.
2. Antonov, V.E., Fig. 11152: System Ce2O3-CeO2 in Phase Equilibria
Diagrams Online, http://ceramics.org/publications-and-resources/
phase-equilibria-diagrams. Accessed 21 March 2011.
3. Zinkevich, M., Djurovic, D., Aldinger, F., Solid State Ionics, 177 (2006)
989–1001.

b1469_Ch-01.indd 40 4/8/2013 12:25:23 PM


b1469 Catalysis by Ceria and Related Materials

Structure, Disorder, Transformation, and Phase Diagram of Ceria-Based Materials 41

4. Kümmerle, E.A., Heger, G., J. Solid State Chem., 147 (1999) 485–500.
5. Yashima, M., Ishimura, D., Yamaguchi, Y., Ohoyama, K., Kawachi, K.,
Chem. Phys. Lett., 372 (2003) 784–787.
6. Yashima, M., Kobayashi, S., Appl. Phys. Lett., 84 (2004) 526–528.
7. Yashima, M., Kobayashi, S., Yasui, T., Solid State Ionics, 177 (2006)
211–215.
8. Knappe, P., Eyring, L., J. Solid State Chem., 58 (1985) 312–314.
9. Ray, S.P., Nowick A.S., Cox., D.E., J. Solid State Chem., 15 (1975)
344–351.
10. Bärnigshausen, H., Schiller, G., J. Less-Common Met., 110 (1985)
385–390.
11. Momma, K., Izumi, F., J. Appl. Crystallogr., 41 (2008) 653–658.
12. Yashima, M., Ishimura, D., Chem. Phys. Lett., 378 (2003) 395–399.
13. Brauer, G., Gingerich, K.A., J. Inorg. Nucl. Chem., 16 (1960) 87–89.
14. Shannon, R.D., Acta Crystallogr., 32 (1976) 751–767.
15. Ozawa, M., Kimura, M., Isogai, A., J. Alloys Compd., 193 (1993) 73–75.
16. Yashima, M., Kakihana, M., Yoshimura, M., Solid State Ionics, 86–88
(1996) 1131–1149.
17. Yashima, M., Mitsuhashi, T., Takashina, H., Kakihana, M., Ikegami, T.,
Yoshimura, M., J. Am. Ceram. Soc., 78 (1995) 2225–2228.
18. Du, Y., Jin, Z., Huang, P., J. Am. Ceram. Soc., 74 (1991) 1569–1577.
19. Yashima, M., Hirose, T., Katano, S., Suzuki, Y., Kakihana, M.,
Yoshimura, M., Phys. Rev. B, 51 (1995) 8018–8025.
20. Yashima, M., Sasaki, S., Kakihana, M., Yamaguchi, Y., Arashi, H.,
Yoshimura, M., Acta Crystallogr. B, 50 (1994) 663–672.
21. Wakita, T., Yashima, M., Acta Crystallogr. B, 63 (2007) 384–389.
22. Yashima, M., Morimoto, K., Ishizawa, N., Yoshimura, M., J. Am. Ceram.
Soc., 76 (1993) 1745–1750.
23. Yashima, M., Arashi, H., Kakihana, M., Yoshimura, M., J. Am. Ceram.
Soc., 77 (1994) 1067–1071.
24. Yashima, M., Sasaki, S., Yamaguchi, Y., Kakihana, M., Yoshimura, M.,
Mori, T., Appl. Phys. Lett., 72 (1998) 182–184.
25. Yashima, M., Morimoto, K., Ishizawa, N., Yoshimura, M., in Science and
Technology of ZIRCONIA V, eds. S.P.S. Badwal, M.J. Bannister R.H.J.
Hannink, Technomic Pub. Co., Lancaster, PA, (1993), pp. 108–116.
26. Tsukuma, K., Shimada, M., J. Mater. Sci., 20 (1984) 1178–1184.

b1469_Ch-01.indd 41 4/8/2013 12:25:23 PM


b1469 Catalysis by Ceria and Related Materials

42 M. Yashima

27. Duwez, P., Odell, F., J. Am. Ceram. Soc., 33 (1950) 274–283.
28. Tani, E., Yoshimura, M., Somiya, S., J. Am. Ceram. Soc., 66 (1983)
506–510.
29. Meriani, S., Mater. Sci. Eng., 71 (1985) 369–370.
30. Meriani, S., J. Phys. (Les Vlis. Fr.), Suppl., 47 (1986) C1-485–489.
31. Meriani, S., Spinolo, G., Powder Diffr., 2 (1987) 255–256.
32. Yashima, M., Wakita, T., Appl. Phys. Lett., 94 (2009) 171902-1–3.
33. Muroi, T., Echigoya, J., Suto, H., Trans. Jpn. Inst. Metals, 29 (1988)
634–641.
34. Meriani, S., Mater. Sci. Eng., A109 (1988) 121–130.
35. Torng, S., Miyazaki, K., Sakuma, T., Ceram. Int., 22 (1996) 309–315.
36. Lamas, D.G., Fuentes, R.O., Fabregas, I.O., Fernandez de Rapp, M.E.,
Lascalea, G.E., Casanova, J.R., Walsoe de Reca, N.E., Craievich, A.F.,
J. Appl. Crystallogr., 38 (2005) 867–873.
37. Larrondo, S., Adelina Vidal, M., Irigoyen, B., Craievich, A.F.,
Lamas, D.G., Fábregas, I.O., Lascalea, G.E., Walsoë de Reca, N.E.,
Amadeo, N., Catal. Today, 107–108 (2005) 53–59.
38. Zhang, F., Chen, C.-H., Hanson, J.C., Robinson, R.D., Herman, I.P.,
Chan, S.-W., J. Am. Ceram. Soc., 89 (2006) 1028–1036.
39. Acuña, L.M., Fuentes, R.O., Lamas, D.G., Fábregas, I.O., Walsöe de
Reca, N.E., Craievich, A.F., Powder Diffraction, 23 (2008) S70–74.
40. Yoshimura, M., Yashima, M., Noma, T., Somiya, S., J. Mater. Sci., 25 (1990)
2011–2016.
41. Yashima, M., J. Phys. Chem. C, 113 (2009) 12658–12662.
42. Yashima, M., Ohtake, K., Kakihana, M., Yoshimura, M., J. Am. Ceram.
Soc., 77 (1994) 2773–2776.
43. Duwez, P., Odell, F., J. Am. Ceram. Soc., 33 (1950) 274–283.
44. Rouanet, A., C. R. Hebd. Seances Acad. Sci., 266 (1968) C-908–911.
45. Longo, V., Roitti, S., Ceramurgia Int., 1 (1971) 4–10.
46. Roitti, S., Longo, V., Ceramurgia Int., 2 (1972) 97–102.
47. Yoshimura, M., Tani, E., Somiya, S., Solid State Ionics, 3–4 (1981) 477–481.
48. Tani, E., Yoshimura, M., Somiya, S., J. Am. Ceram. Soc., 66 (1983)
506–510.
49. Duran, P., Gonzalez, M., Moure, C., Jurado, J.R., Pascual, C., J. Mater.
Sci., 25 (1990) 5001–5006.

b1469_Ch-01.indd 42 4/8/2013 12:25:23 PM


b1469 Catalysis by Ceria and Related Materials

Structure, Disorder, Transformation, and Phase Diagram of Ceria-Based Materials 43

50. Yashima, M., Takashina, H., Kakihana, M., Yoshimura, M., J. Am.
Ceram. Soc., 77 (1994) 1869–1874.
51. Du, Y., Yashima, M., Koura, T., Kakihana, M., Yoshimura, M., Script
Metall. Mater., 31 (1994) 327–332.
52. Kuznctsov, F.A., Rezukhina, F.N., Zhur. Fiz. Khim., 34 (1960)
C-2467–2468.
53. Westrum Jr., E.F., Beale Jr., A.F., J. Phys. Chem., 65 (1961) 353–355.
54. Ricken, M., Nolting, J., Riess, I., J. Solid State Chem., 54 (1984) 89–99.
55. Gallagher, S.A., Dworzak, W.R., J. Am. Ceram. Soc., 68 (1985) C-206–207.
56. Yashima, M., Morimoto, K., Ishizawa, N., Yoshimura, M., J. Am. Ceram.
Soc., 76 (1993) 2865–2868.
57. Yashima, M., Yoshimura, M., Jpn. J. Appl. Phys. 31 (1992) L1614–1617.
58. Várez, A., Jolly, J., Oliete, P., Sanjuán, M.L., Garcia-Gonzalez, E.,
Jardiel, T., Sanz, J., Inorg. Chem., 48 (2009) 9693–9699.
59. Kresse, G., Joubert, D., Hafner, J., Phys. Rev. B, 59 (1999) 1758–1775.
60. Rodriguez, J.A., Hanson, J.C., Kim, J.-Y., Liu, G., Iglesias-Juez, A.,
Fernandez-Garcia, M., J. Phys. Chem. B, 107 (2003) 3535–3543.
61. Conesa, J., J. Phys. Chem. B, 107 (2003) 8840–8853.
62. Tibiletti, D., Amieiro-Fonseca, A., Burch, R., Chen, Y., Fisher, J.M.,
Goguet, A., Hardacre, C., Hu, P., Thompsett, D., J. Phys. Chem. B, 109
(2005) 22553–22559.
63. Yang, Z., Fu, Z., Wei, Y., Lu, Z., J. Phys. Chem. C, 112 (2008)
15341–15347.
64. Yang, Z., Fu, Z., Wei, Y., Hermansson, K., Chem. Phys. Lett., 450 (2008)
286–291.
65. Nolan, M., Fearon, J.E., Watson, G.W., Solid State Ionics, 177 (2006)
3069–3074.
66. Castleton, C.W., Kullgren, J., Hermansson, K., J. Chem. Phys., 127
(2007) 244704-1–11.
67. Da Silva, J.L.F., Verónica Ganduglia-Pirovano, M., Sauer, J., Bayer, V.,
Kresse, G., Phys. Rev. B, 75 (2007) 045121-1–10.
68. Andersson, D.A., Simak, S.I., Skorodumova, N.V., Abrikosov, I.A.,
Johansson, B., Appl. Phys. Lett., 90 (2007) 031909-1–3.
69. Jomard, G., Petit, T., Pasturel, A., Magaud, L., Kresse, G., Hafner, J.,
Phys. Rev. B, 59 (1999) 4044–4052.

b1469_Ch-01.indd 43 4/8/2013 12:25:23 PM


b1469 Catalysis by Ceria and Related Materials

44 M. Yashima

70. Kuwabara, A., Tohei, T., Yamamoto, T., Tanaka, I., Phys. Rev. B, 71
(2005) 064301-1–7.
71. Jaffe, J.E., Bachorz, R.A., Gutowski, M., Phys. Rev. B, 72 (2005)
144107-1–9.
72. Igawa, N., Ishii, Y., J. Am. Ceram. Soc., 84 (2001) 1169–1171.
73. Yashima, M., Tsunekawa, S., Acta Crystallogr. B, 61 (2006) 161–164.
74. Yashima, M., Oh-uchi, K., Tanaka, M., Ida, T., J. Am. Ceram. Soc., 89
(2006) 1395–1399.
75. Yashima, M., Solid State Ionics, 179 (2008) 797–803.
76. Yashima, M., J. Ceram. Soc. Jpn., 117 (2009) 1055–1059.
77. Yashima, M., Itoh, M., Inaguma, Y., Morii, Y., J. Am. Chem. Soc., 127
(2005) 3491–3495.
78. Yashima, M., Sirikanda, N., Ishihara, T., J. Am. Chem. Soc., 132 (2010)
2385–2392.
79. Yashima, M., Yonehara, Y., Fujimori, H., J. Phys. Chem. C, 115 (2011)
25077–25087.
80. Chen, Y.-C., Yashima, M., Ohta, T., Ohoyama, K., Yamamoto, S.,
J. Phys. Chem. C, 116 (2012) 5246–5254.
81. Yashima, M., Takizawa, T., J. Phys. Chem. C, 114 (2010) 2385–2392.
82. Duprez, D., Descorme, C., in Catalysis by Ceria and Related Materials, ed.
A. Trovarelli, Imperial College Press, London, (2002), pp. 243–280.
83. Wakita, T., Yashima, M., Appl. Phys. Lett., 92 (2008) 101921-1–3.
84. Boaro, M., Leitenburg, C., Dolcetti, G., Trovarelli, A., J. Catal., 193
(2000) 338–347.
85. Faber Jr., J., Seitz, M.A., Mueller, M.H., J. Phys. Chem. Solids, 37 (1976)
903–907.
86. Faber Jr., J., Seitz, M.A., Mueller, M.H., J. Phys. Chem. Solids, 37 (1976)
909–915.
87. Faber Jr., J., Physica B, 150 (1988) 241–249.
88. Berber, K., Martin, U., Mursic, Z., Schneider, J., Boysen, H., Frey, F.,
Mater. Sci. Forum, 79–82 (1991) 685–690.
89. Yashima, M., Kobayashi, S., Yasui, T., Faraday Discussions, 134 (2007)
369–376.
90. Wang, D.Y., Park, D.S., Griffth, J., Nowick, A.S., Solid State Ionics, 2
(1981) 95–105.

b1469_Ch-01.indd 44 4/8/2013 12:25:23 PM


b1469 Catalysis by Ceria and Related Materials

Structure, Disorder, Transformation, and Phase Diagram of Ceria-Based Materials 45

91. Sammes, N.M., Tompsett, G.A., Näfe, H., Aldinger, F., J. Eur. Ceram.
Soc., 19 (1999) 1801–1826.
92. Battle, P.D., Catlow, R.A., Drennan, J., Murray, A.D., J. Phys. C: Solid
State Phys., 16 (1983) L561–566.
93. Yashima, M., Ishimura, D., Appl. Phys. Lett., 87 (2005) 221909-1–3.
94. Battle, P.D., Catlow, C.R.A., Heap, J.W., Moroney, L.M., J. Solid State
Chem., 63 (1986) 8–15.
95. Battle, P.D., Catlow, C.R.A., Heap, J.W., Moroney, L M., J. Solid State
Chem., 67 (1987) 42–50.
96. Lerch, M., Boysen, H., Radaelli, P.G., J. Phys. Chem. Solids, 58 (1997)
1557–1568.
97. Eguchi, K., Setoguchi, T., Inoue, T., Arai, H., Solid State Ionics, 52
(1992) 165–172.
98. Otsuka-Yao, S., Morikawa, H., Izu, N., Okuda, K., J. Jpn. Inst. Metals, 59
(1995) 1237–1246.
99. Thomson, J.B., Armstrong, A.R., Bruce, P.G., J. Am. Chem. Soc., 118
(1999) 11129–11133.
100. Otsuka-Yao-Matsuo, S., Izu, N., Omata, T., Ikeda, K., J. Electrochem. Soc.,
145 (1998) 1406–1413.
101. Omata, T., Kishimoto, H., Otsuka-Yao-Matsuo, S., Ohtori, N.,
Umesaki, N., J. Solid State Chem., 147 (1999) 573–583.
102. Thomson, J.B., Armstrong, A.R., Bruce, P.G., J. Solid State Chem., 148
(1999) 56–62.
103. Sasaki, T., Ukyo, Y., Kuroda, K., Arai, S., Muto, S., Saka, H., J. Ceram.
Soc. Jpn., 111 (2003) 382–385.
104. Sasaki, T., Ukyo, Y., Suda, A., Sugiura, M., Kuroda, K., Arai, S.,
Saka, H., J. Ceram. Soc. Jpn., 112 (2004) 440–444.
105. Achary, S.N., Sali, S.K., Kulkarni, N.K., Krishna, P.S.R., Shinde, A.B.,
Tyagi, A.K., Chem. Mater., 21 (2009) 5848–5859.

b1469_Ch-01.indd 45 4/8/2013 12:25:23 PM


b1469 Catalysis by Ceria and Related Materials

This page intentionally left blank

b1469_Ch-01.indd 46 4/8/2013 12:25:23 PM


b1469 Catalysis by Ceria and Related Materials

CHAPTER 2

UNDERSTANDING CERIA-BASED
CATALYTIC MATERIALS:
AN OVERVIEW OF RECENT PROGRESS
Juan José Delgado, Eloy del Río, Xiaowei Chen,
Ginesa Blanco, José María Pintado, Serafín Bernal
and José Juan Calvino
Departamento de Ciencia de los Materiales e Ingeniería Metalúrgica y
Química Inorgánica, Facultad de Ciencias, Universidad de Cádiz,
11510-Cádiz, Spain

2.1 Introduction
The literature that has appeared since the publication of the first
edition of Catalysis by Ceria and Related Materials,1 some ten years ago,
has allowed us to identify a number of major breakthroughs in the
chemical and nanostructural characterization of this highly inter-
esting family of catalytic materials, which are discussed in the differ-
ent sections of this chapter.
Among the driving forces responsible for these recent develop-
ments, hydrogen production for fuel cells is certainly one of them.
To illustrate this trend with actual figures, we could simply mention
that a literature search comprising simultaneously the keywords
hydrogen production, catalyst and fuel cells returns currently nearly
1,500 references within the period 2002–2011, whereas this number
drops dramatically to just 85 in the much larger 1900–2001 time
span. From the references corresponding to the most recent period
over 200 contain ceria or ceria–zirconia as keywords. In addition to

47

b1469_Ch-02.indd 47 4/8/2013 12:26:11 PM


b1469 Catalysis by Ceria and Related Materials

48 J. J. Delgado et al.

this, the recent literature shows that other relevant applications of


ceria and ceria–zirconia catalysts are for the abatement of pollutant
emissions from both stationary and mobile sources and, less inten-
sively, for the production of chemicals. In most cases, the redox
properties of the ceria-based materials are considered to play a key
role in determining their catalytic behaviour.
Steam and dry reforming of hydrocarbons2–13 (mainly methane)
and oxygenates;14–17 the use of methane/steam redox cycles;18–20
methane partial oxidation;21–27 CO oxidation;28–33 preferential oxida-
tion of CO in the presence of large amounts of hydrogen (PROX)34,35
and low-temperature water gas shift (LT-WGS)36–47 are major pro-
cesses investigated for hydrogen production. Some of these reac-
tions are in fact those which were classically investigated in
connection with three-way catalysis (TWC), unquestionably the most
successful technological application of ceria-based catalysts. This
may possibly explain the movement of the research community
working in TWCs to the study of fuel cells.
Concerning depollution strategies, the investigation of Pd-only
TWCs is possibly one of the hot topics in the most recent litera-
ture,48–60 together with the study of total oxidation processes: hydro-
carbon combustion,61–65 selective soot particulate removal,66–78 volatile
organic compound (VOC) elimination79–81 and catalytic wet-air oxida-
tion of organics (CWAO)82–84 and NOx abatement.85–93
Finally, for the production of chemicals, ceria-based catalysts have
been applied to processes such as the conversion of syngas into alco-
hols,94 C–C coupling reactions,95 partial oxidation reactions,96,97 direct
synthesis of dimethyl carbonate,98,99 esterification,100 reactions involving
aldolic condensations101,102 and selective hydrogenation.103–125
To connect the previous paragraphs with the matter addressed
in this chapter, in summary we will consider the following two
points: (1) the redox nature of most of these reactions, and (2) the
participation in many cases of CO as either reactant or product.
A second subject worthy of comment is the sudden arrival of
supported gold catalysts. Haruta’s reports,126–130 in the late 1980s,
about the exceptional catalytic activity of gold when adequately pre-
pared in the form of nanometre-sized particles not only broke a

b1469_Ch-02.indd 48 4/8/2013 12:26:12 PM


b1469 Catalysis by Ceria and Related Materials

Understanding Ceria-Based Catalytic Materials 49

‘well’-established dogma in heterogeneous catalysis but also pro-


vided possibly the clearest illustration of the influence of nano-
structure on chemical properties. Following the initial impulse of
Haruta’s work, research into catalysis using gold has experienced
continuous growth, the period from 2002 to 2011 being particularly
productive. To illustrate this with some figures, there has been an
increase in one order of magnitude, from about 200 to over 2,000,
in the number of papers on gold catalysts between the periods
1900–2001 and 2002–2011. Therefore we could label the period
between the two editions of the present book as a ‘golden’ time.
Finally, a third event key to understanding the present chapter
is the extraordinary technological advancements experienced in the
last decade in electron microscopy instrumentation and methodolo-
gies. First, the implementation of so-called electron lens correctors
has solved the spatial resolution limitations imposed by intrinsic
electron-optical aberrations (e.g. spherical aberration) both in
transmission and scanning transmission operation modes.131–138 The
incorporation in the microscope column of extra elements (correc-
tors), which compensate for the unavoidable aberrations of the
round condenser and objective lenses has led to imaging with sub-
angstrom resolution and chemical analysis at the level of atomic
columns.139–143 Thus, in the latest generation of electron micro-
scopes, those equipped with this new technology, it is possible to
record not only images, both in bright-field high-resolution trans-
mission electron microscopy (HRTEM) and dark-field modes of
high-angle annular dark-field scanning transmission electron
microscopy (HAADF-STEM), in which a representation of the
atomic structure of solid materials with 0.08-nm resolution is attain-
able, but also to perform chemical analysis (in some cases including
the oxidation state of the elements) on single atomic columns.
There are another two technological advances in high-resolution
electron microscopy: (1) the development of electron gun mono-
chromators, which have reduced the energy resolution attainable
in spectroscopic techniques (electron energy loss spectroscopy
(EELS), X-ray energy dispersive spectroscopy (XEDS) to below 100
meV and (2) the recording and processing of angularly resolved

b1469_Ch-02.indd 49 4/8/2013 12:26:12 PM


b1469 Catalysis by Ceria and Related Materials

50 J. J. Delgado et al.

information, which has allowed a transition from 2D to 3D


tomography.144–155
The structural analysis of catalytic materials has strongly bene-
fited from these developments138,156–162 not only because the resolu-
tion limits have now been set at the atomic scale, but also because
they have made it possible to analyse simultaneously a structure
using different complementary techniques. Apart from an improved
spatial and energy resolution, aberration correction has also mini-
mized the so-called contrast delocalization effects, which in
conventional uncorrected imaging, complicate the structural
analysis of surfaces and interfaces at an atomic scale, an issue of
great interest in catalysis research. The absence of delocalization
together with a morphological description, which is now possible
using 3D shape reconstruction from electron tomography, pro-
vides very relevant information about surfaces and interfaces in
catalytic materials.
All these aspects–the redox properties of ceria–zirconia catalysts,
supported gold catalysts and the application of advanced electron
microscopy techniques–are the essential ingredients of this chapter,
which will be organized as follows. In Section 2.2 there is an analysis
of the relation between the nanostructure and redox properties of
bare ceria–zirconia oxides after high-temperature redox aging. The
whole set of results included in this section will allow us to establish
a genuine model that correlates thermal treatment, nanostructure
and the macroscopic redox response. Section 2.3 will then deal with
gold-supported ceria–zirconia catalysts and the analysis will concen-
trate on the H2 and CO adsorption capacity of this type of catalyst.
We will develop, by putting together pieces of information obtained
at the atomic and macroscopic scales, a very detailed model of how
the adsorption of these two probe molecules takes place on gold
nanoparticles and the ceria–zirconia support.
In addition to the particular conclusions that will be drawn for
each of these topics, the two sections are well suited to the general
theme of the chapter, emphasizing the message that an in-depth
nanostructural analysis gives an understanding of macroscopic
performance.

b1469_Ch-02.indd 50 4/8/2013 12:26:12 PM


b1469 Catalysis by Ceria and Related Materials

Understanding Ceria-Based Catalytic Materials 51

2.2 A Model for the Nanostructure, Pre-Treatment


Conditions and Redox Behaviour in High-Temperature
Aged Ceria–Zirconia Catalysts
As already mentioned, the oxygen handling capability of ceria–zirconia
mixed oxides is a key factor in their catalytic applications. Actually,
most of the processes listed in the introduction imply oxygen exchange
processes with the reactants. This explains the ongoing interest in the
investigation of the redox properties of these materials.

2.2.1 Oxygen storage capacity of CeO2–ZrO2 vs CeO2


Early reports on the textural and chemical effects of adding ZrO2 to
pure CeO2 not only evidenced a definite improvement in resistance
against sintering at high temperature compared to pure ceria,163–165
but also a significant increase in the capacity to release lattice oxy-
gen under reducing conditions; a property usually referred to as
oxygen storage capacity (OSC) and quantified on the basis of either
conventional H2 temperature-programmed reduction (TPR) or oxy-
gen uptake experiments. The first reports of this chemical result166,167
date back to 1990, as cited in a paper by Murota et al.168 in which a
comparative analysis of H2-TPR experiments carried out on CeO2–
ZrO2 materials with a zirconia content ranging from 0 mol% up to
93 mol% is made, see Fig. 2.1(a).
Further, more detailed, work by Kašpar’s group not only con-
firmed the larger low-temperature reducibility of the metal-loaded
(Rh, Pt) Ce–Zr mixed-oxide catalysts, but also evidenced the benefi-
cial effect of their improved reducibility in NO reduction by CO.169
Likewise, Ozawa et al. observed improvements in CO, NOx and hydro-
carbon removal activity, under dynamic air–fuel ratio conditions, after
the addition of a Ce1−xZrxO2/Al2O3 redox promoter to a Pt/Al2O3
catalyst.170 Using oxides prepared by alternative methods, which
allow the production of high-surface-area materials, similar findings
have been reported.171 Further analysis on the origin of the changes
in OSC after the addition of ZrO2 in metal-loaded samples, led
Kašpar et al. to propose a direct link between the increase in the

b1469_Ch-02.indd 51 4/8/2013 12:26:12 PM


b1469 Catalysis by Ceria and Related Materials

52 J. J. Delgado et al.

1
1 1

MS signal m/c=18 (a.u.)


H2 Consumption(a.u.)

H2 Consumption(a.u.)
2 2
3
3 2
4
5
3 6
4
7
4 8
5 9
5 10
6
200 400 600 800 1000 400 600 800 1000 1200 400 600 800 1000 1200
Temperature (K) Temperature (K) Temperature (K)
(a) (b) (c)

Figure 2.1 (a) TPR curves of zirconium-containing CeO2: (1) CeO2, (2) 90CeO2–
10ZrO2, (3) 70CeO2–30ZrO2, (4) 50CeO2–50ZrO2, (5) 15CeO2–85ZrO2, (6) 7CeO2–
93ZrO2; (b) TPR profiles of: (1) fresh Ce0.5Zr0.5O2 and (2–3) recycled 1 and 7 times,
respectively, (4) fresh CeO2 and (5) CeO2 recycled 2 times; (c) Influence of the pre-
treatment conditions on the reducibility of a Ce/Zr mixed oxide (CZ-68/32-LS).
TPR-MS traces corresponding to the sample pre-treated as follows: (1) fresh oxide
sample, (2) SR pre-treatment, (3) SR-SO, (4) SR-MO, (5) SR-MO, (6) SR-SO, (7)
SR-MO, (8) SR-MO, (9) SR-SO, (10) SR-MO (abbreviations are explained in the text).
Experiments (2)–(6) were recorded successively on the same oxide sample. Adapted
from Murota et al.,168 and Baker et al.179 with permission from Elsevier, and adapted
from Fornasiero et al.180 with permission from the Royal Society of Chemistry.

degree of bulk reduction and an improvement in oxygen mobility in


the mixed oxides,172,173 which in turn is due to structural distortion of
the M–O bonds (mainly affecting Zr–O), detected upon formation of
solid solutions.174,175 Using conductivity measurements and the Nernst–
Einstein relation, Trovarelli et al.176 found that in the 573–1173 K
temperature range the oxygen diffusion coefficients for Ce–Zr mixed
oxides were two orders of magnitude larger than those determined
for pure ceria, which they correlated with an increase in the CO oxi-
dation activity of the former when the reaction was carried out in
transient pulse mode or by cycling the feedstream composition.
Therefore, these studies show a direct correlation between OSC
and the bulk structure of the mixed oxides. It is nevertheless also

b1469_Ch-02.indd 52 4/8/2013 12:26:12 PM


b1469 Catalysis by Ceria and Related Materials

Understanding Ceria-Based Catalytic Materials 53

worth mentioning that more detailed studies of both static177 and


dynamic178 OSC values determined in metal-loaded (Rh, Pd, Pt)
ceria–zirconia catalysts have evidenced the eventual contribution to
these quantities of surface-related processes, such as hydrogen
spillover, in those experiments where H2 is used as reducing agent.
Likewise, it was also shown that when CO is used as reductant a com-
plicated dynamic CO-OSC behaviour is observed, with surface
reduction, CO storage and CO desorption being detected as pro-
cesses contributing to the OSC values.
These two papers illustrate how detailed experimentation is
actually required to fully understand the oxygen-handling behaviour
of metal-loaded ceria–zirconia materials, and how far this behaviour
could be from correlations of a simplistic bulk structure to redox
performance. Such a naïve approach disregards the subtle nuances
of the chemical interactions, which can occur between different
chemical species (H2 or CO) and this type of system, and it totally
fails to explain the influence of structural and compositional param-
eters, as well as those of the experimental measurement conditions,
on the observed OSC values. We will keep this in mind hereafter.

2.2.2 Effect on OSC of high-temperature redox cycles


In addition to the increase of reducibility at low temperatures, the
most remarkable feature in the redox behaviour of the bare oxides
was the improvement observed in the OSC after aging under reduc-
tion/oxidation cycles at high temperature. Thus, Fornasiero et al.180
showed, Fig. 2.1(b), that on treating a Ce0.5Zr0.5O2 oxide in 5% H2/
Ar at 1273 K and further re-oxidizing it at 700 K, a shift of the
H2-TPR peaks towards lower temperatures takes place. Despite the
oxide sintering induced by this treatment, the temperature of
reduction in a subsequent H2-TPR run was found to have decreased
from 900 K to 700 K.181 Moreover, the improved redox response per-
sisted after a high number of successive reduction/re-oxidation
cycles as mentioned above.
This behaviour is the opposite of that observed in pure
ceria,182,183 for which the same thermochemical treatment leads to

b1469_Ch-02.indd 53 4/8/2013 12:26:13 PM


b1469 Catalysis by Ceria and Related Materials

54 J. J. Delgado et al.

the disappearance of the low-temperature reduction peak, i.e. to a


clear deterioration of its reducibility. This particular observation is
very relevant because it shows that the redox response of ceria–
zirconia mixed oxides subjected to high-temperature aging treat-
ments is much better than that of ceria, a critically important feature
in TWC applications.
Further studies on this issue have revealed some additional, very
important, details. Thus, Baker et al.179 proved, using H2-TPR meas-
urements monitored by mass spectrometry, that the redox response
of mixed oxides after a high-temperature reduction treatment
(severe reduction or SR) strongly depended on the temperature at
which the final re-oxidation step was carried out. Thus, if after
reduction at a high temperature (5% H2/Ar at 1223 K), re-oxidation
was performed at a low temperature (5% O2/He at 823 K, mild oxi-
dation or MO), a displacement of the H2-TPR trace towards lower
temperatures was observed. If re-oxidation was performed at a high
temperature (5% O2/He at 1223 K, severe oxidation or SO), no
significant improvement could be observed with respect to the non-
cycled, fresh, oxide; see Fig. 2.1(c). Taking these results into consid-
eration, it was also proved that the state of the mixed oxide could be
reversibly switched between that corresponding to an oxide with
improved redox performance, i.e. with low-temperature features in
the H2-TPR trace, and that corresponding to a conventional, unim-
proved, ceria–zirconia oxide. In effect, as shown in Fig. 2.1(c), an SR
followed by an MO treatment takes the fresh oxide to the improved
redox response state, as can be seen by comparison of traces (1) and
(2) in the figure. If this SR-MO oxide is further treated using the SO
routine, the H2-TPR trace, (3) in Fig. 2.1(c), shifts back to higher
temperatures and becomes quite similar to that of the original or
fresh oxide. This pattern is reproduced, see traces (4)–(10) in
Fig. 2.1(c), in further, successive, SR-MO, SR-SO cycles. Izu et al. also
reported this reversible redox response, for ceria–zirconia oxides of
different compositions, on the basis of oxygen evolution measure-
ments performed using an electrochemical cell.184
Given the practical relevance of this effect, much research effort
has been devoted to clarifying its origin as well as its relation to the

b1469_Ch-02.indd 54 4/8/2013 12:26:13 PM


b1469 Catalysis by Ceria and Related Materials

Understanding Ceria-Based Catalytic Materials 55

structure of the mixed oxides. It is worth mentioning first that these


questions have been scrutinized in detail both at the macroscopic
and microscopic, or even better atomic, levels by using a large variety
of both chemical and structural techniques. Thus, comparative
textural, structural, spectroscopic as well as theoretical, analyses of
ceria–zirconia oxides in different states of the redox cycles, at both
bulk and surface levels, have been undertaken in the most recent
literature by means of N2 adsorption at 77 K, X-ray diffraction
(XRD), extended X-ray absorption fine structure spectroscopy
(EXAFS), X-ray absorption near-edge spectroscopy (XANES),
Raman spectroscopy, electron and neutron diffraction, X-ray photo-
electron spectroscopy (XPS), XEDS, EELS, HRTEM, HAADF-STEM,
electron tomography, OSC measurements, H2 and CO chemisorp-
tion, or DFT, among others. But even with such a huge variety of
information and the wide perspective gained, unravelling these
questions has proved to be a rather complex and very challenging
task, which has caused strong controversy and remained an open
issue until very recently.185–187 The large structural and compositional
flexibility of the CeO2–ZrO2 system, in which a high number of both
stable and metastable phases have been reported188,189 (cubic, differ-
ent tetragonal variants and monoclinic) and, most importantly, the
highly spatially localized nature of the features which influence the
redox response and their extreme sensitivity to the thermochemical
story of the oxide, easily explain these difficulties.
We should mention that assembling all the pieces in the puzzle
has required information from techniques, which have just become
fully available only recently, and also a detailed analysis of the chemi-
cal response of the oxides by very carefully designed experiments
and, finally, reasoning far beyond a simple correlation between the
bulk structure of the oxides and their redox response.

2.2.3 Chemical and structural investigation of OSC changes


after high-temperature redox cycling
From the above, it is fairly clear that a high temperature (> 1173 K)
reducing treatment is the triggering step that moves the fresh oxide

b1469_Ch-02.indd 55 4/8/2013 12:26:13 PM


b1469 Catalysis by Ceria and Related Materials

56 J. J. Delgado et al.

into a state in which an improved redox response can be obtained,


if a proper, low-temperature, re-oxidation treatment is applied.
Structural analysis using Rietveld-fitted X-ray powder diffracto-
grams,184,187,190,191 Raman spectroscopy186,187,191 and electron micros-
copy192–194 have shown that by starting from a fresh ceria–zirconia
sample typically exhibiting a random distribution in its cationic
sublattice, the high-temperature reduction treatment-induced
ordering in the Ce and Zr distribution, produces an oxide, Ce2Zr2O7,
with a pyrochlore-like structure. If this heavily reduced sample is re-
oxidized at moderate temperatures (Treoxn ≤ 873 K), the cationic
sublattice remains unaltered as in pyrochlore, thus resulting in an
ordered oxidized, Ce2Zr2O8, phase.190,191 By contrast, if it is re-oxi-
dized at a high temperature (Treoxn ≥ 1173 K), its cationic sublattice
rearranges back to a random distribution.187,191
The electron microscopy results in Fig. 2.2 illustrate this idea.
Figures 2.2(a)–(c) show representative HRTEM images of (a) a fresh,
(b) an SR-MO and (c) an SR-SO Ce0.62Zr0.38O2 oxide. Fourier analysis
of the image contrasts indicates that the image of the fresh catalyst
can be interpreted as due to a Ce–Zr mixed-oxide solid solution with
a cubic structure viewed along the [110] zone axis (see López-Haro
et al.66 for a detailed discussion of how the different phases reported
in the ceria–zirconia phase diagram can be detected by electron
microscopy techniques); spots corresponding to the 0.312 nm {111}

3 nm

[110] [001]

111

(1-11)

002 (-111) (002)

(a) (b) (c)

Figure 2.2 HRTEM images for a Ce0.62Zr0.38O2 oxide after the following treatments:
(a) fresh oxide; (b) SR-MO; (c) SR-SO. Adapted from López-Haro et al.66 with per-
mission from John Wiley & Sons.

b1469_Ch-02.indd 56 4/8/2013 12:26:13 PM


b1469 Catalysis by Ceria and Related Materials

Understanding Ceria-Based Catalytic Materials 57

and 0.270 nm {200} type planes can be measured on the digital dif-
fraction pattern (DDP) shown as the inset in Fig. 2.2(a). After
SR-MO, the image contrast does not change dramatically, but the
DDPs clearly show the appearance of superstructure reflections,
marked on the inset in Fig. 2.2(b). These extra superstructure spots
are located at ½ {111} and ½ {200}, which is in good agreement with
the transformation of the solid solution with cubic fluorite structure
into the Ce2Zr2O8 pyrochlore. This last structural type has a unit cell
whose dimensions are double those of fluorite due to ordering, in an
alternating fashion, of Ce and Zr species along the three major <100>
cubic directions. Finally, after SR-SO, both the HRTEM images and
DDP show only features characteristic of a simple fluorite-type cell.
Bearing in mind that superstructure-type HRTEM image contrasts
may have their origin either in changes related to the cationic sublat-
tice or in the oxygen sublattice (e.g. by modulation of the oxygen
displacements or oxygen vacancy ordering), as suggested in Torng
et al.195 and Masui et al.,196 the interpretation of the HRTEM studies
proposed in the paragraphs above,192 suggesting the occurrence of a
disorder-order transition in the cationic sublattice of the mixed oxide,
could only be confirmed after a subsequent HAADF-STEM study.197
Image contrasts in HAADF-STEM images depend on, roughly,
the squared average value of the atomic number (Z 2) in each atomic
column.198 Under these conditions, in a material like the mixed
oxides we are considering, which contain oxygen (Z = 8) and heavy
metal elements, Zr (Z = 40) and Ce (Z = 58), image contrasts feature
mostly the location of the latter. Thus, white contrasts observed in
these images correspond, for those projections in which the cations
and anions do not overlap, to the position of Ce and Zr species.
Likewise, higher intensities would be expected at sites populated by
Ce species. With this simple image interpretation, a comparison of
the HAADF-STEM images in Fig. 2.3 clearly reveals significant
changes in the arrangement of cations in the mixed oxides between
the fresh or SR-SO and the SR-MO materials. In the former,
Fig. 2.3(a), the DDPs show reflections corresponding to a fluorite cell
whereas in the SR-MO oxide, Fig. 2.3(b), the ½-type reflections are
clearly visible, confirming the presence of a superstructure, which is

b1469_Ch-02.indd 57 4/8/2013 12:26:13 PM


b1469 Catalysis by Ceria and Related Materials

58 J. J. Delgado et al.

Ce

Zr Ce/Zr

(a) (b)

Figure 2.3 HAADF-STEM images for a Ce0.62Zr0.38O2 oxide after the following
treatments: (a) SR-SO; (b) SR-MO. Adapted from López-Haro et al.66 with permis-
sion from John Wiley & Sons.

necessarily due, given the origin of the image contrasts, to an


ordered arrangement of the cation sublattice. A more detailed dis-
cussion of the interpretation of HRTEM and HAADF-STEM images
of ceria–zirconia nano-crystals as well as an in-depth analysis of the
complementarity of these two techniques can be found in a recent
review by López-Haro et al.66
Recently, López-Haro and co-workers199 have analysed the struc-
ture of the SR-MO state of a Ce0.5Zr0.5O2 oxide using atomically
resolved chemical imaging. In this study, the authors definitely con-
firmed, by direct chemical analyses performed atomic column by
atomic column, the ordering of the cationic sublattice into a
pyrochlore-like arrangement. They produced images that show the
differences in chemical composition of adjacent atomic columns
expected for a pyrochlore-like superstructure. In fact, detailed analy-
ses of the electron energy loss spectra recorded in each atomic col-
umn have allowed the authors to conclude, by precise comparison
with simulations, that after the reduction treatment employed to
prepare the SR-MO sample they studied, 5 h in flow of pure H2 at
1223 K, cation ordering was not completed. Local deviations from

b1469_Ch-02.indd 58 4/8/2013 12:26:13 PM


b1469 Catalysis by Ceria and Related Materials

Understanding Ceria-Based Catalytic Materials 59

the stoichiometry were detected at subnanometric level, which


appeared in the form of a residual fractional population by Zr of unit
cell positions which in the exact pyrochlore structure are pure Ce.
These very interesting data clearly indicate a certain compositional
flexibility of the pyrochlore structure and the inherent difficulties in
producing well-defined compositions of the Ce–Zr–O system. Most
of the time intermediate, metastable situations are reached since the
applied thermochemical treatments are not severe or prolonged
enough to allow for a true equilibrium final state.
In summary, the electron microscopy data indicate that under
severe reduction conditions, which lead to an oxide with a large
Ce3+ content, the solid-solution structure becomes unstable and a
disorder-order reaction takes place that transforms the fluorite cell
into a pyrochlore one. This pyrochlore phase is therefore the stable
one under high-temperature reducing conditions.
Conversely, as a number of authors have stressed,200,201 under
oxidizing conditions, the pyrochlore-related phase is thermody-
namically unstable against the transition leading to an oxide with a
disordered distribution in its cationic sublattice and finally to segre-
gated ceria- and zirconia-rich phases.202,203 Therefore, the prepara-
tion of an oxidized ordered phase can only be achieved by mild
re-oxidation of the heavily reduced oxide (i.e. under conditions at
which the rate of the order-disorder process is negligible), as with
the MO process. The fully oxidized SR-MO oxide is thus a metasta-
ble phase. Accordingly, if it is heated at high enough temperatures,
under an oxygen partial pressure allowing the oxide to be kept in a
fully oxidized state, a new phase showing a random distribution of
Ce and Zr in the cationic sublattice is formed.204,205
Taking all this nanostructural information into account, it
becomes tempting to establish a direct correlation between the
improved redox performance and the presence of a pyrochlore-type
structure in the bulk of the metastable SR-MO oxide. Thus, on the
basis of DFT studies in which the tetragonal and pyrochlore phases
of a Ce0.5Zr0.5O2 mixed oxide were compared, Wang et al. recently
proposed that the difference in OSC between the two structures is
due to a higher localization of the structural relaxation effects

b1469_Ch-02.indd 59 4/8/2013 12:26:14 PM


b1469 Catalysis by Ceria and Related Materials

60 J. J. Delgado et al.

around oxygen vacancies in the pyrochlore-like Ce2Zr2O8 phase.206


Though this argument could perfectly explain the intrinsic thermo-
dynamic difference in reducibility between the two structures, it
does not give a full interpretation of all the evidence available for
the redox behaviour of the high-temperature aged mixed oxides in
the whole experimental range of investigated temperatures.
As we will see, the differences in OSC are not solely influenced
by thermodynamic factors. Depending on the temperature at which
the OSC is measured and also on the nature of the reducing agent
(H2, CO), other surface-related processes may have a significant
influence on the measured OSC values. Kinetic factors and, in con-
nection with them, the surface structure of the oxides have a rele-
vant role. In general, the interplay between surface and bulk
structure determines the observed total OSC values. As with the
OSC differences between ceria and ceria–zirconia metal-loaded
catalysts, a more in-depth analysis is required in order to fully under-
stand the redox response of thermally aged oxides.
The experiments reported in Yeste et al.207 definitely contributed
in clarifying the limitations imposed by kinetic factors for low-
temperature OSC measurements. In this paper Yeste et al. subjected
a low-surface-area sample of Ce0.62Zr0.38O2 to an SR treatment at 1223
K for 5 h under pure H2 flowing at 500 cm3.min−1. A portion of this
SR-treated oxide was then oxidized under mild conditions, 5% O2/
He at 773 K for 1 h, and a second one under severe conditions, 5%
O2/He at 1123 K for 5 h. The H2-TPR of these two samples shows the
distinct features already described for SR-MO and SR-SO samples,
traces (A) and (B) in Fig. 2.4(a), i.e. a displacement to higher tem-
peratures of the water evolution peaks for the SR-SO oxide.
Nevertheless, when a small amount of rhodium (0.3 wt%) was sup-
ported on both oxides, the H2-TPR diagrams of the two metal-oxide
systems are closer, traces (C) and (D) in Fig. 2.4(a), their character-
istic peak featuring at much lower and closer temperatures than in
the case of the corresponding bare oxides. As proposed in Yeste
et al.207 the presence of the dispersed Rh phase facilitates H2 dissocia-
tion, the resulting atomic H being further transferred onto the sup-
port via a spillover mechanism. Once hydrogen is dissociated at a

b1469_Ch-02.indd 60 4/8/2013 12:26:14 PM


b1469 Catalysis by Ceria and Related Materials

Understanding Ceria-Based Catalytic Materials 61

4
m/e: 18
SR SO m/e: 2
SR SO H1123
SR MO
A 3

Intensity (a.u.)
Intensity (a.u.)

B
2

1
C

400 600 800 1000 1200 350 400 450 500 550 600 400 600 800 1000 1200
Temperature (K) Temperature (K) Temperature (K)
(a) (b) (c)

Figure 2.4 (a) Influence of supported rhodium on the redox response of ceria–
zirconia mixed oxides. Temperature-programmed reaction mass spectroscopic
(TPR-MS) study. Traces for the H2O (m/e (mass/charge ratio) = 18) signal for
SR-MO (A), SR-SO (B), 0.3% Rh/SR-MO (C), and 0.3% Rh/SR-SO (D) samples.
Trace B’ was recorded after applying the first series of ultimate OSC measurements
to the SR-SO sample. (b) Volumetric study of hydrogen chemisorption on
Ce0.62Zr0.38O2 samples. Initial hydrogen pressure was 38 torr. The samples were sub-
mitted to successive cycles of heating at the indicated temperatures for 30 min fol-
lowed by cooling to 298 K always under hydrogen. Pressure drops were determined
at 298 K . SR-SO (first series) accounts for the sample resulting from the first series
of ultimate OSC measurements carried out on the SR-SO sample. (c) TPR-MS
traces for H2 (m/e = 2) and H2O (m/e = 18) corresponding to a Ce0.62Zr0.38O2 sam-
ple submitted, respectively, to SR-MO and SR-SO aging routines. To facilitate the
comparison between H2 and H2O traces, hydrogen consumption is plotted as a
positive signal. The effect of supported Rh on the reducibility of the oxides is exem-
plified by the diagrams recorded for the 0.3% Rh/SR-SO sample, bottom part of
the figure. Adapted from Yeste el al.207,208 with permission from the American
Chemical Society, and adapted from Bernal et al.212 with permission from Elsevier.

low temperature, the difference in reducibility observed between


the two bare oxide samples becomes less strong, which suggests that
the different abilities of the oxides to activate hydrogen dissociation
has a significant influence on their low-temperature reducibility, in
the absence of the metal phase.

b1469_Ch-02.indd 61 4/8/2013 12:26:14 PM


b1469 Catalysis by Ceria and Related Materials

62 J. J. Delgado et al.

Following the reasoning in the paragraph above, the dissociation of


H2 must be faster on the SR-MO sample than on the SR-SO one. This
was actually confirmed by H2 chemisorption studies carried out on the
bare oxides, Fig. 2.4(b).208 If we recall that earlier studies on H2 adsorp-
tion for ceria209,210 and ceria–zirconia211 proved that the process implies
the dissociation of the molecule with inherent formation of atomic
chemisorbed species, the data in this figure suggest that, over the whole
range of studied temperatures, the number of H atoms/nm2 is signifi-
cantly higher in the SR-MO oxide. In particular, in the 373–473 K
temperature range, the surface density of chemisorbed hydrogen is
approximately six times larger for the SR-MO sample. We may conclude
therefore that H2 activation is much faster for the SR-MO samples.
The kinetic nature of the limitations observed in H2 adsorption
for the SR-SO sample was further confirmed by extending the heat-
ing/cooling cycles under H2 up to 573 K, a temperature still below
that at which the reduction of the SR-SO sample starts, see trace C
in Fig. 2.4(a). As shown in Fig. 2.4(b), for SR-SO, the amount of
chemisorbed hydrogen progressively increases with temperature,
the value recorded at 573 K being even higher than that determined
at 473 K for SR-MO. These observations strongly suggest the
relevance of the H2 chemisorption step in the control of the low-
temperature reducibility of the mixed-oxide samples.
The comparison of the m/e = 18 (water) and m/e = 2 (H2)
signals in H2-TPR experiments followed by mass spectrometry,
Fig. 2.4(c), indicates a parallel with no delay, evolution of water and
H2 consumption,212 which may also be considered as a clear indica-
tion of the role of hydrogen adsorption in the control of the overall
low-temperature redox response of the samples.
Ultimate, or total, OSC data measured after reduction in H2 at
increasing temperatures,208 gathered in Table 2.1 and also plotted in
Fig. 2.5, allow us to make comments on the influence of kinetic fac-
tors. If it is assumed that the ultimate OSC values determined in the
Rh-loaded systems represent the thermodynamic limit of reducibility,
it is clear that, as suggested in Wang et al.,206 this thermodynamic
reducibility of the SR-MO oxide is higher than that of the SR-SO one
over the whole temperature range, even at 1173 K, a temperature at
which a significant difference in OSC between the two oxides is still

b1469_Ch-02.indd 62 4/8/2013 12:26:14 PM


b1469 Catalysis by Ceria and Related Materials

Understanding Ceria-Based Catalytic Materials 63

Table 2.1 Ultimate H2-OSC values corresponding to bare and 0.3%


Rh-loaded SR-MO and SR-SO oxide samples.
T (K) SR-MO 0.3% Rh/SR-MO SR-SO 0.3% Rh/SR-SO
473 8 45 0 23
623 45 55 2 31
773 63 65 22 39
973 74 75 55 —
1173 82 82 70 67

100

CZ-MO/H2
90
0.3%Rh/CZ-MO/H2
CZ-SO/H2
80
0.3%Rh/CZ-SO/H2
70

60
% Ce3+

50

40

30

20

10

0
373 473 573 673 773 873 973 1073 1173 1273
Temperature (K)

Figure 2.5 Evolution with reduction temperature of the ultimate H2-OSC values
determined for bare and rhodium-loaded (0.3 wt%) SR-SO and SR-MO oxides.
Adapted from Yeste et al.208 with permission from the American Chemical Society.

observed. Nevertheless, the OSC values determined for the bare


oxides, in the low-temperature region, do not reach the thermody-
namic limit of reducibility, as they are clearly controlled by kinetic
factors. As marked on Fig. 2.5, for the SR-MO sample the

b1469_Ch-02.indd 63 4/8/2013 12:26:14 PM


b1469 Catalysis by Ceria and Related Materials

64 J. J. Delgado et al.

convergence of the kinetically controlled OSC values with those


determined for the corresponding Rh-loaded sample (thermody-
namic limit) occurs at approximately 773 K, and at an even higher
temperature for the SR-SO sample.
The question that rises now is how the surface structure changes
between the fresh or SR-SO samples, which show a similar redox
performance, on the one hand, and the SR-MO oxide, with enhanced
reducibility, on the other. Advanced electron microscopy techniques
have proved to be unique tools to answer this relevant issue.
Thus, as reported in Hernández et al.,197 the SR-MO treatment
induces, in addition to the already noted disorder-order transition
in the bulk of the oxide, which transforms the cubic or tetragonal
Ce–Zr solid solution into a pyrochlore-related phase, significant tex-
tural and compositional changes on the surface.
In particular, studies by HAADF-STEM electron tomography
have clearly revealed a definite increase in the extent of {111} type
facets exposed at the surface. From a situation in which a balanced
population of both rounded and faceted crystallites are observed in
a fresh or SR-SO sample, the morphology of most of the SR-MO
crystallites features shapes close to octahedrons or other more
irregular shapes, in which {111} facets are still dominant, Fig. 2.6.66

(a) (b) (c)

(d)

Figure 2.6 Electron tomography reconstructions corresponding to a SR-MO


Ce0.62Zr0.38O2 oxide. (a) Octahedron-like nano-crystallites whereas (b) shows crystal-
lites with more irregular shapes which still expose (c) and (d) mostly {111} type
facets. Reproduced from López-Haro et al.66 with permission from John Wiley & Sons.

b1469_Ch-02.indd 64 4/8/2013 12:26:14 PM


b1469 Catalysis by Ceria and Related Materials

Understanding Ceria-Based Catalytic Materials 65

More importantly, a detailed analysis of atomically resolved


HAADF-STEM images using image simulation techniques deter-
mined the structural and compositional information about the first
{111} atomic planes exposed just at the surface of the octahedron-
like SR-MO crystallites. We will explain this important point in more
detail.
As depicted in Fig. 2.7(a), when the structure of a fully oxidized
Ce2Zr2O8 pyrochlore is projected along any direction perpendicular
to the <111> direction, the distribution of the cation sublattice can
be easily described in terms of a stacking of {111} planes. Two vari-
ants of these {111} planes are present: a Zr-rich one and a Ce-rich

Zr rich variant
Ce rich variant

a b c

surface
d

Figure 2.7 (a) Supercell of a Ce2Zr2O8 pyrochlore-like structure projected


through the [110] direction for a 4.8-nm thick crystal and a 7° tilt along [1–11].
(b) Experimental HAADF-STEM image of SR-MO oxide. (c) HAADF-STEM simu-
lated image corresponding to the model in (a). (d) and (e) intensity profiles along
a direction perpendicular to the surface of the experimental and simulated images,
respectively. Solid red arrows indicate Ce-rich planes; dashed blue arrows indicate
Zr-rich planes. Adapted from Hernández et al.197 with permission from the
American Chemical Society.

b1469_Ch-02.indd 65 4/8/2013 12:26:15 PM


b1469 Catalysis by Ceria and Related Materials

66 J. J. Delgado et al.

one. For a 50/50 pyrochlore the Ce/Zr molar ratio in the Ce-rich
planes is 75/25 whereas it is 25/75 in the Zr-rich variant. Piling up
these two variants, in an alternating fashion, along the <111> direc-
tion gives rise to the final structure.
Due to the direct relation existing between contrast intensity
and atomic number, HAADF-STEM images of the pyrochlore struc-
ture along these directions show lines of alternating high and low
intensity, see Fig. 2.7(b). The high-intensity lines indicate the posi-
tions of the Ce-rich {111} planes, whereas the low-intensity ones cor-
respond to the Zr-rich variant. The intensity profiles from the
experimental images clearly reveal, as expected, an alternation of
the intensities, which, according to a simulated image, Fig. 2.7(c),
can be interpreted as proposed.
From the intensity profiles, Figs. 2.7(d) and (e), the stacking
sequence can be continuously tracked from the bulk of the crystal-
lites outwards to the surface. Moreover, from an observation of the
image intensity level at the surface, a conclusion can be drawn about
which kind of {111} variant is exposed. In this respect, the intensity
profiles recorded from the experimental images, Fig. 2.7(d), repre-
sentative of an SR-MO oxide, indicate the presence at the surface of
a low-intensity Zr-rich plane.
Therefore, electron microscopy data have revealed that the
disorder-order transformation taking place in the bulk is accompa-
nied by parallel structural and compositional rearrangements at the
surface. The {111} facets resulting from the bulk transformation,
which are Zr rich and cation ordered, provide more effective routes
for dissociating H2.
In Yeste et al.208 the key influence of the surface structure on the
OSC has been explored in more detail. The evolution of total OSC
with reduction temperature was studied for a Ce0.62Zr0.38O2 sample
submitted to two additional treatments: (1) SR-MO followed by a
short, 30 minute, oxidation at medium temperature, 1073 K, under
5% O2/He (SR-MO-O1073) and (2) SR-SO followed by a short, 1 h,
reduction at 1123 K under 5% H2/Ar (SR-SO-H1123). OSC data for
these samples are presented in Fig. 2.8, together with those for
SR-MO and SR-SO oxides.

b1469_Ch-02.indd 66 4/8/2013 12:26:15 PM


b1469 Catalysis by Ceria and Related Materials

Understanding Ceria-Based Catalytic Materials 67

Figure 2.8 Evolution with reduction temperature of the ultimate H2-OSC values
determined for a Ce0.62Zr0.38O2 oxide after the treatments indicated in the leg-
ends. Adapted from Yeste et al.208 with permission from the American Chemical
Society.

Note how, in the SR-MO-O1073 sample, the application of a short


oxidation treatment to the SR-MO sample causes a deterioration in
the low-temperature redox behaviour, whereas the high-temperature
OSC values remain unaltered. Something similar, but opposite, hap-
pens for the SR-SO-H1123 sample. In this case the low-temperature
redox response increases but, again, the high-temperature OSC values
remain close to those of the SR-SO oxide.
These results suggest that the short oxidation treatment
applied to the SR-MO oxide transforms it, with regard to the low-
temperature redox response, into the SR-SO type but it keeps it in
the SR-MO state for the high-temperature response. Similarly, the
SR-SO-H1123 oxide shows an SR-MO type of low-temperature
redox behaviour but, that is characteristic of SR-SO at high
temperatures.
Electron microscopy studies performed by HRTEM and STEM-
XEDS have provided the necessary data to rationalize these

b1469_Ch-02.indd 67 4/8/2013 12:26:15 PM


b1469 Catalysis by Ceria and Related Materials

68 J. J. Delgado et al.

(a) (d)

(b) (c)

Figure 2.9 HRTEM images for a Ce0.68Zr0.32O2 oxide: (a) SR-MO-O1073;


(b) SR-MO-O1123; (c) SR-MO-O1123; (d) SR-SO-H1123. Adapted from Yeste et al.208
with permission from the American Chemical Society.

observations. Thus, as shown in Fig. 2.9(a), for the SR-MO-O1073


oxide, HRTEM images show crystallites that still have a pyrochlore-
like structure in the bulk. DDPs obtained from these images clearly
show the presence of the ½-type superstructure spots, inset in
Fig. 2.9(a). Nevertheless the surface of these crystallites has clearly
changed, appearing now much rougher and more rounded.
Occasionally, small globular nodules with a solid-solution-type struc-
ture have been detected at the surface, as illustrated in Fig. 2.9(c).
For the SR-SO-H1123 oxide, HRTEM images show in general
crystallites with the solid-solution-type bulk structure, but small
patches close to the surface of the crystallites can now be detected,
Fig. 2.9(d), in which the structure of the cation-ordered pyrochlore
is evident.

b1469_Ch-02.indd 68 4/8/2013 12:26:16 PM


b1469 Catalysis by Ceria and Related Materials

Understanding Ceria-Based Catalytic Materials 69

SR-SO

SR-MO-O1073

SR-MO

Figure 2.10 XEDS study of Ce–Zr distribution in SR-MO, SR-SO and SR-MO-O1073
oxide nano-crystals. Data expressed as atomic percentage of Ce: 100[Ce/(Ce + Zr)].
Adapted from Yeste et al.208 with permission from the American Chemical Society.

In summary, these HRTEM results show that the short redox treat-
ments produce subtle changes at the surface of SR-MO and SR-SO
oxides but they are not strong enough to modify the bulk structure.
The STEM-XEDS study shown in Fig. 2.10 provides some addi-
tional interesting pieces of information regarding the short oxida-
tion treatments. This figure summarizes the nano-analytical data
corresponding to a series of XEDS studies carried out on SR-MO,
SR-SO and SR-MO-O1073 samples. Remarkable differences may be
noticed from one oxide to another. In Fig. 2.10 it can be seen that
SR-MO has the narrowest compositional distribution, with most of
the analysed nano-regions exhibiting Ce/Zr molar ratios close to
the nominal value, 62:38. By contrast, the sample resulting from the
high-temperature oxidizing aging treatment, SR-SO, has the broad-
est range of Ce/Zr values.
This observation suggests compositional segregation, a process
that is known to be induced in oxides with Ce/Zr molar ratios typi-
cally ranging from 50:50 to 70:30, by thermal aging (T > 1173 K),

b1469_Ch-02.indd 69 4/8/2013 12:26:16 PM


b1469 Catalysis by Ceria and Related Materials

70 J. J. Delgado et al.

under oxidizing conditions.202,203 In accordance with the X-ray dif-


fraction study reported in Colon et al.,202 the segregation process
leads to the formation of two new phases that are indexed as ceria–
zirconia solid solutions with approximate Ce/Zr molar ratios of
80:20 and 20:80. However, as also noted in this reference, for an
oxide with Ce/Zr 50:50, a prolonged (140 h) calcination at 1473 K
was required to complete the process. For Ce/Zr 62:38, even higher
temperatures must be applied. Thus, very likely, an SR-SO oxide,
aged under oxygen at 1223 K for 5 h, does not represent a true equi-
librium state for the system but an intermediate transient on the way
to the final fully segregated state.
The wideness of the compositional distribution determined for
SR-MO-O1073, although closer to that of SR-MO, is between those
for SR-MO and SR-SO, as shown in Fig. 2.10. We should conclude,
accordingly, that SR-MO-O1073 might well correspond to the very
first stages of the process of transforming the SR-MO oxide (i.e. the
sample mainly consisting of an ordered Ce–Zr sublattice, with a high
compositional homogeneity) into two Ce- and Zr-rich segregated
phases.
These characterization data allow us to draw two very important
conclusions: (1) the short oxidation and reduction treatments lead
to an initial, first stage, transformation of the SR-MO and SR-SO
oxides; (2) these first stages of the transformation start and affect
primarily the surface structure of the oxides while, to a large extent,
their bulk features are retained.
These electron microscopy results allow a proper rationalization
of the OSC values shown in Fig. 2.8 and highlight once more the key
influence of the surface on low-temperature OSC measurements. In
effect, the decrease in the low-temperature OSC observed for the
SR-MO-O1073 sample is due to the deterioration of the pyrochlore-
type surface features observed by HRTEM, whereas the increase in
low-temperature OSC in the SR-SO-H1123 oxide is due to the
appearance at the surface of small nuclei with pyrochlore structure.
Likewise, since high-temperature OSC is mainly dominated by ther-
modynamic factors and the short redox treatments mostly do not
modify the bulk structure of the starting oxides, the high-temperature

b1469_Ch-02.indd 70 4/8/2013 12:26:16 PM


b1469 Catalysis by Ceria and Related Materials

Understanding Ceria-Based Catalytic Materials 71

SR-MO
SR-SO-H1123
SR-MO-O1123
SR-SO

Figure 2.11 Volumetric study of H2 chemisorption for SR-MO, SR-MO-O1123,


SR-SO and SR-SO-H1223. The experiments consisted of heating the samples under
H2 (PH2 = 38 torr) for 30 min at each of the indicated temperatures, followed by cooling
to 298 K under H2. H2 adsorption was determined from the pressure drop at 298 K.
Adapted from Yeste et al.208 with permission from the American Chemical Society.

OSC values observed for SR-MO-O1073 and SR-SO-H1123 coincide


with those of the SR-MO and SR-SO oxides, respectively.
The hydrogen adsorption data shown in Fig. 2.11 corroborate
further the definite influence of the short redox treatments on
the surface structure of the SR-MO and SR-SO oxides as well as the
interpretation proposed above. Thus, note how the low-temperature
hydrogen adsorption capacity of the SR-MO oxide clearly decreases
after the short oxidation treatment, whereas after a short reduction
of the SR-SO oxide, the hydrogen adsorption capacity of this oxide
clearly increases.
As shown in Fig. 2.8, the transformation of the redox response
of the SR-MO oxide into that of the SR-SO oxide proceeds continu-
ously with increasing severity of the oxidation treatment. In fact, the
OSC values determined for a SR-MO-O1123 sample, i.e. a SR-MO

b1469_Ch-02.indd 71 4/8/2013 12:26:16 PM


b1469 Catalysis by Ceria and Related Materials

72 J. J. Delgado et al.

Figure 2.12 Scheme depicting the pretreatment nanostructure redox behaviour


relations.

sample submitted to a short, slightly more severe, oxidation treat-


ment, lie between those for SR-MO-O1073 and SR-SO, linked to a
structural transformation in which the surface of the SR-MO oxide
has been modified to a larger extent, Fig. 2.9(c).
The whole set of chemical and structural data discussed so far can
be put together into a consistent model that allows a rationalization
of the modifications, with thermochemical treatments, of the ultimate
H2-OSC values of ceria–zirconia oxides. This model takes into account
both the surface and bulk structure of the oxides and the influence of
both thermodynamic and kinetic factors and is shown in Fig. 2.12.
The picture stresses the closed loop between the different situations
we have described in this section and the nature of the thermochemi-
cal treatments necessary to proceed from one situation to another.
This comprehensive model, which explains all the experimental
data currently available for the H2-OSC behaviour of bare oxides,
can be considered a fruitful result of a general experimental

b1469_Ch-02.indd 72 4/8/2013 12:26:16 PM


b1469 Catalysis by Ceria and Related Materials

Understanding Ceria-Based Catalytic Materials 73

approach combining precisely designed chemical measurements


and a battery of electron-microscopy-based nanostructural charac-
terisation techniques.

2.2.4 CO-OSC of ceria–zirconia bare oxides


There is some quite interesting data from recent work by Bernal and
co-workers213 for the OSC behaviour of these oxides when CO is
used as reductant instead of hydrogen.
The first interesting point we should mention in connection
with the reducibility of the mixed oxides under CO is that behav-
iour similar to that observed with H2 can be observed for the high-
temperature redox aged oxides. Figure 2.13 shows the results of
differential thermo-gravimetric (DTG) experiments run under both
CO and H2 for SR-MO and SR-SO oxides. Note that, qualitatively,
the behaviour of the two oxides is the same under hydrogen and

CO-OSC H2-OSC

SR-SO
SR-SO

SR-MO
SR-MO

(a) (b)

Figure 2.13 DTG diagrams corresponding to (a) the reduction in flow of 5% CO/
He of SR-SO and SR-MO oxide samples, heating rate 10 K.min−1; (b) the reduction
in flow of 5% H2/Ar of SR-SO and SR-MO oxide samples, heating rate 10 K.min−1.
Prior to the experiments the oxides were in all cases heated under flowing 5% O2/
He at 773 K (1 h). Adapted from Yeste et al.213 with permission from Elsevier.

b1469_Ch-02.indd 73 4/8/2013 12:26:17 PM


b1469 Catalysis by Ceria and Related Materials

74 J. J. Delgado et al.

CO. The SR-MO oxide has weight loss peaks at lower temperatures
than the SR-SO one. Therefore the reducibility of SR-MO under CO
seems higher than that of the SR-SO oxide.
Compared to the corresponding H2-DTG traces, the main reduc-
tion peak in the CO-DTG diagrams is significantly shifted towards
lower temperatures. We should conclude, accordingly, that for both
oxide samples the low-temperature reducibility under flowing 5%
CO/He is higher than that observed in a flow of 5% H2/Ar. From a
thermodynamic point of view, for temperatures below 900 K, the
reducing power of CO is superior to that of H2, the difference pro-
gressively increasing as the temperature decreases. Accordingly, in
the low-temperature range, CO should be expected to behave as a
more effective reductant than H2.
Nevertheless, the question arises again whether the recorded
CO-DTG traces are mainly determined by thermodynamic factors or
if kinetic aspects of the process must also be considered in the inter-
pretation of these diagrams. An analysis of the ultimate OSC data,
shown in Table 2.2 and Fig. 2.14, will shed some light on this issue.
As deduced from Table 2.2 and Fig. 2.14, both the H2-OSC and
CO-OSC data show that, throughout the whole range of investigated
temperatures, the degree of reduction by the SR-MO sample is
higher than for SR-SO. This is a consequence of the change in the
relative thermodynamic stability of the ordered and disordered struc-
tures as a function of the redox state of the samples, Fig. 2.15(a).
Under oxidizing conditions, the higher stability corresponds to the

Table 2.2 Ultimate H2- and CO-OSC values for bare and 0.3% rhodium-loaded
SR-MO and SR-SO oxide samples.
SR-MO/ SR-MO/ 0.3% Rh/ SR-SO/ SR-SO/ 0.3% Rh/
T (K) CO H2 SR-MO/H2 CO H2 SR-SO/H2
473 28 8 45 5 0 23
623 59 45 55 27 2 31
773 72 63 65 47 22 39
973 79 74 75 59 55
1173 83 82 82 69 70 67

b1469_Ch-02.indd 74 4/8/2013 12:26:17 PM


b1469 Catalysis by Ceria and Related Materials

Understanding Ceria-Based Catalytic Materials 75

Figure 2.14 Evolution with reduction temperature of the ultimate H2-OSC and
CO-OSC values determined for the systems indicated in the legend. Adapted from
Yeste et al.213 with permission from Elsevier.

phase having a disordered arrangement in the cationic sublattice,


i.e. to the SR-SO sample, whereas the opposite is true for oxides in
their reduced state.
The comparative analysis of the CO-OSC and H2-OSC data
recorded for SR-MO and SR-SO also deserves some comments. For
the same oxide sample, in the low-temperature range (T < 773 K),
the CO-OSC values are significantly larger than those of H2-OSC. By
contrast, for T > 973 K, much closer values are experimentally
observed. This is in qualitative agreement with the variation of ∆G0r
for CO and H2 oxidation reactions against T. As shown in Fig.
2.15(b), the slope of the approximately linear plot of ∆G0r vs. T is
higher for the former reaction, the crossing point occurring at
approximately 900 K. For T < 900 K, the relative reducing power of
CO progressively increases with respect to that of H2, as the

b1469_Ch-02.indd 75 4/8/2013 12:26:17 PM


b1469 Catalysis by Ceria and Related Materials

76 J. J. Delgado et al.

temperature decreases. In contrast, above 973 K, H2 is slightly more


effective. If this effect is combined with the evolution of the thermo-
dynamic parameters for the reduction of the oxides with their redox
state,205 Fig. 2.15(a), we may understand the general trends followed
by the CO-OSC and H2-OSC data as the temperature is increased.
As shown in Fig. 2.14, the largest differences between CO-OSC
and H2-OSC data are found at T < 773 K. If we look at this region of
Fig. 2.14, some remarkable observations can be made. As already
discussed, for kinetic reasons, in a flow of H2, the low-temperature
reducibility of the oxides is strongly enhanced by the presence of
small amounts of Rh supported on them. Assuming that the Rh
phase does not modify the thermodynamic properties of the oxides,
the dramatic influence of the metal should be interpreted as due to
a catalytic effect. On rhodium, H2 adsorption is faster, thus favouring
the subsequent transfer of atomic hydrogen to the support via spill-
over. If so, the H2-OSC value determined for the Rh-containing
samples would actually provide oxygen storage data much closer to
the thermodynamic limit of the oxide reduction than those for bare
oxides, at the same temperature. This is a relevant observation.
If the H2-OSC data reported in Table 2.2 for Rh/SR-MO and Rh/
SR-SO are compared with those of CO-OSC for the corresponding

py-Ce2Zr2O8
∆G0

c,t-Ce0.5Zr0.5O2

cc,t-Ce0.5Zr0.5O1.75
c,

py-Ce2Zr2O7

0 100 %Ce3+

(a) (b)

Figure 2.15 (a) The expected change in the relative thermodynamic stability of
the fully oxidized and fully reduced states of the pyrochlore-like and solid-solution-
type phases of a Ce0.5Zr0.5O2 oxide. (b) ∆Gor vs. T plots for CO and H2 oxidation
reactions. Adapted from Yeste et al.213 with permission from Elsevier.

b1469_Ch-02.indd 76 4/8/2013 12:26:17 PM


b1469 Catalysis by Ceria and Related Materials

Understanding Ceria-Based Catalytic Materials 77

SR-MO

SR-SO

Figure 2.16 Ultimate CO-OSC study. DTG traces corresponding to the stepwise
reduction of SR-MO (full line) and SR-SO (dotted line) in a flow of 5% CO/He.
Heating rate between successive steps: 10 K.min−1. Duration of the isothermal steps:
1 h. Prior to the experiments the oxides were heated under flowing 5% O2/He at
773 K (1 h). Adapted from Yeste et al.213 with permission from Elsevier.

bare oxides, we may notice that, at 473 K, the former are much larger
than the latter. As already discussed, this is unexpected from the
thermodynamic point of view. Actually, in accordance with Fig. 2.15,
if a true equilibrium state had been reached, at 473 K, the CO-OSC
data should be larger than the corresponding H2-OSC values.
Therefore, as for the H2-OSC values, for the bare oxides the
CO-OSC data recorded at the lowest temperatures are determined
by kinetic rather than thermodynamic factors. This conclusion is
also supported by the shape of the stepwise CO-DTG diagrams, as
shown in Fig. 2.16. As clearly shown in the first step of the CO-DTG
trace for SR-MO in Fig. 2.16, after 1 h isothermal heating at 473 K,
the rate of reduction is still measurable, i.e. the equilibrium reduced
state was not reached at the end of this isothermal step. The same is
true for SR-SO even at 623 K, which suggests that the kinetic restric-
tions are, as already happened for H2, even stronger for this sample.

b1469_Ch-02.indd 77 4/8/2013 12:26:17 PM


b1469 Catalysis by Ceria and Related Materials

78 J. J. Delgado et al.

The latter proposal is also supported by the remarkable upwards


shift of the main peak in the continuous CO-DTG trace for SR-SO
compared to that for SR-MO, Fig. 2.13.
Additional studies are still necessary to give a full understanding
of the origin of the observed kinetic control in the ultimate CO-OSC
measurements. It should be noted, however, that, in accordance
with the results reported in Table 2.2 for the oxide-supported
Rh samples, the oxygen diffusion step is unlikely to be the rate-
controlling step of the overall reduction process under flowing CO.
Therefore, the kinetic control is probably associated with either the
CO adsorption step leading to the formation of carbonate species214
or to the subsequent decomposition of the above mentioned
carbonates. Currently available studies on the mechanism of CO
oxidation over ceria–zirconia under dynamic conditions214,215 sug-
gest that the decomposition of the surface carbonate species could
play a role in the kinetics of ceria–zirconia reduction by CO.
Nevertheless, the eventual contribution of CO adsorption to the
rate-controlling step should not be disregarded.
These final comments connect perfectly with the topic addressed
in the next section: the interaction of CO with ceria–zirconia-supported
gold catalysts.

2.3 Chemical Characterization of Cerium-Based


Oxide-Supported Gold Catalysts
As outlined in the introduction to this chapter, this section will
briefly review the chemical characterization studies on cerium-based
oxide-supported noble metal (NM) catalysts that have been pub-
lished since the first edition of Catalysis by Ceria and Related Materials1
in 2002. We suggest a reading of this book for a detailed and critical
account of earlier studies on the topic.
In recent years, the most significant progress has been made in the
chemical characterization of cerium-based oxide-supported gold cata-
lysts. There are a number of reasons that justify this observation. First,
the main principles governing the chemical behaviour of the classic
NM(Rh, Pd, Pt)/CeO2 and ceria-related catalysts were reasonably well

b1469_Ch-02.indd 78 4/8/2013 12:26:17 PM


b1469 Catalysis by Ceria and Related Materials

Understanding Ceria-Based Catalytic Materials 79

established before 2002 (Bernal et al.216 and references therein). A sec-


ond major reason is the discovery in 1987 of the extremely high CO
oxidation activity of gold nanoparticles dispersed on suitable 3D metal-
oxide supports.126 This initial finding, which at that time was totally
unexpected, can be seen some twenty years later as one of the major
scientific events in the recent history of heterogeneous catalysis.217 In
fact, an exponential growth of research activity on oxide-supported
gold catalysts followed Haruta’s pioneering work.217
The adsorption of reactants is known to be a key step in hetero-
geneous catalysis. Accordingly, detailed qualitative and quantitative
information about the surface chemistry of gold is of outmost impor-
tance for a fine understanding of its exceptional and very puzzling
catalytic behaviour. With reference to the catalysts of supported noble
metals from Groups 8–10, for which chemical characterization rou-
tines are reasonably well established, the development of appropriate
tools for providing useful chemical information on highly dispersed
gold catalysts is still a very challenging issue. The extreme nobility of
gold218 is a major reason for this. Thus, it is generally acknowledged
that gold has much lower surface activity than the metals of Groups
8–10 against H2 and CO, two probe molecules commonly used in the
chemical characterization of noble metal catalysts. In fact, until some
twenty years ago the surface chemistry of gold was almost completely
unknown.218 Because of the series of reviews published successively
over the last 20 years,122,218–223 however, the amount and quality of
information available on the surface chemistry of gold has grown
very rapidly. At present, adsorption studies include a variety of
probe molecules and atoms,220 those dealing with CO,218–221,223–256
H2,104,114,122,219–221,233,241,253,256–270 and O2218,220,221,233,271–281 being particularly
numerous. Regarding the structural nature of the investigated gold
samples, studies on massive single crystals, polycrystalline thin films,
foils, wires and powders,220,221,227,271,282,283 as well as clusters and nano-
particles dispersed on planar model220,229,230,232,236,242,243,245–247,254,284,285 and
powder223–236,231,234,235,239,241,248–253,256,260,261,270,286–291 oxide supports are pres-
ently available. Some studies on inverse oxide/Au model systems have
also been reported.268,292 In parallel with the experimental papers or in
combination with them, the chemisorptive properties of both neutral

b1469_Ch-02.indd 79 4/8/2013 12:26:17 PM


b1469 Catalysis by Ceria and Related Materials

80 J. J. Delgado et al.

and charged clusters, small nanoparticles, and even oxide-supported


gold systems, have been increasingly investigated from a theoretical
point of view. In most cases, different DFT approaches were used in
these calculations.104,220,222,233,236,237,255,257,258,262,265,266,277–279,283,291,293–308
In this section, recent progress in the chemical characteriza-
tion of conventional cerium-containing oxide-supported gold pow-
der catalysts will be briefly discussed. We will focus our attention
on H2 and CO chemisorption studies. There are several reasons
justifying this choice. As discussed in Bernal et al.216 H2 and CO are
by far the most commonly used probe molecules in the character-
ization of noble metals supported on ceria and closely related
mixed oxides. Also very importantly, ceria-based gold catalysts
are known to be highly active materials for CO oxida-
tion,218,220,239,272,273,309–328 LT-WGS,39–42,240,309,329–356 selective oxidation
of CO in the presence of H2 (PROX)244,253,310,312,326,333,335,357–361 or
selective hydrogenation103,105,106,108,124 reactions in which these two
molecules are involved. Therefore, to gain detailed information
about the chemistry of H2- and CO-catalyst interactions, it is crucial
to arrive at a deeper understanding of these reactions, whose
mechanisms have not been fully interpreted as yet.
As discussed in Bernal et al.,216 cerium-containing oxide-
supported noble metal catalysts have a number of specificities,
which makes their chemical characterization particularly challeng-
ing. First, ceria-based oxides may adsorb large amounts of H2 and
CO.216 If so, to separate metal and support contributions to the total
amount of the adsorbed probe molecules is mandatory in order to
fully interpret the corresponding adsorption results. Moreover, as
has been well established for several NM/CeO2(CeO2–ZrO2) sys-
tems (NM: Rh, Pd, Pt),216 and more recently for analogous gold
catalysts,241,251,267,270 the interaction of both H2 and CO molecules
with ceria and closely related oxides is strongly enhanced by the
presence of a supported metal phase. This represents an additional
serious complication because it prevents the straightforward use of
parallel adsorption studies on the bare supports as a tool for deter-
mining their contribution to the total adsorption by the metal cata-
lysts. Second, because of the reducible nature of cerium-based

b1469_Ch-02.indd 80 4/8/2013 12:26:17 PM


b1469 Catalysis by Ceria and Related Materials

Understanding Ceria-Based Catalytic Materials 81

oxides,216 the corresponding supported metal catalyst may exhibit a


strong metal/support interaction (SMSI) effect, like that first
reported by Tauster et al. in 1976.362 This issue is discussed at length
in Bernal et al.216,363 where a model accounting for the specific char-
acteristics of this phenomenon in NM/CeO2(CeO2–ZrO2) (NM:
Rh, Pd, Pt) catalysts and their differences from those of classic NM/
TiO2 systems are analysed. As is well known, the occurrence of
strong metal/support interaction effects implies that the chem-
isorptive properties of the metal phase against H2 and CO may be
reversibly modified by the redox state of the support, a significant
deactivation of the metal phase being observed on catalysts pre-
reduced at an appropriate temperature. In this respect, it is inter-
esting to outline that earlier studies364,365 on Rh and Pt supported on
the same ceria–zirconia mixed oxide have shown that the specific
characteristics of the strong metal/support interaction effects are
sensitive to the nature of the metal phase. It is therefore of interest
to investigate the behaviour of gold with reference to that exhibited
by the above mentioned noble metals. Moreover, the eventual
occurrence of such an effect in gold catalysts could be important to
fully interpret their behaviour in processes like PROX, LT-WGS or
the selective hydrogenation reactions mentioned above, that typi-
cally occur under net reducing conditions.
This section has been organized in two subsections. The first,
Section 2.3.1, gives a brief overview of the latest studies dealing with
the interaction of H2 with supported gold catalysts, and more spe-
cifically with Au/CeO2 and closely related systems. In the second,
Section 2.3.2, we will analyse the latest data on the use of CO as a
probe molecule for characterizing the family of cerium-containing
oxide-supported gold catalysts. As will be discussed below, some very
recent studies have proved that the appropriate combination of
Fourier transform infrared (FTIR) spectroscopy and volumetric
adsorption studies of CO, with electron microscopy, computer simu-
lation and nanostructural modelling techniques, provides a wealth
of very useful chemical information about this family of cerium-
containing oxide-supported gold powder catalysts. Particularly
noticeable are the correlations that have been established between

b1469_Ch-02.indd 81 4/8/2013 12:26:17 PM


b1469 Catalysis by Ceria and Related Materials

82 J. J. Delgado et al.

chemical properties, as revealed by CO adsorption studies at room


temperature, on the one hand, and the nanostructural constitution,
or redox state of the support, on the other hand.

2.3.1 Hydrogen adsorption on cerium-based oxide-supported gold


catalysts
2.3.1.1 Preliminary studies on hydrogen adsorption on Au single-crystal
and thin-film surfaces
Compared to the metals of Groups 8–10, gold is acknowledged to
have a rather limited ability to chemisorb H2.116,122,218,219,257,260
Experimental studies have shown that molecular H2 is not dissoci-
ated on a clean Au(110)(1 × 2) surface; atomic hydrogen generated
by a hot filament is required for observing adsorbed H species.366 As
revealed by thermal desorption (TD), at a heating rate of 15 K.s−1,
the H2 desorption peak occurs at a temperature as low as 216 K, thus
indicating that recombination takes place rather easily on this
single-crystal surface.366 The activation energy (Ea) of the recombina-
tion process has also been determined by TD. Assuming a
second-order reaction, a pre-exponential factor in the range of 10−1
to 10−3 cm2.s−1 and a coverage of 0.5 ML, the estimated Ea was found
to be 51 kJ.mol−1.366 These experimental observations are consistent
with DFT calculations,257,262 in accordance with which no dissociation
of H2 should be expected to occur on Au{111} and Au{001} surfaces.
In effect, as reported in Corma et al.,262 the calculated energies for
the dissociation of H2 on both model surfaces are clearly endother-
mic. Actually, the theoretical calculations predict that the only very
slightly exothermic adsorption processes on these single-crystal sur-
faces are those characterized by H2-surface complexes in which the
optimized energies of adsorption, H–H and H-surface distances were
found to be < 2 kJ.mol−1, 0.075 nm and 0.4 nm, respectively.262 These
data clearly support the occurrence of an extremely weak molecular
interaction between H2 and the Au{111} and Au{001} surfaces.
In contrast to the surface behaviour of single crystals, dissocia-
tive adsorption of H2 (D2) may occur on thin films of gold.367,368

b1469_Ch-02.indd 82 4/8/2013 12:26:17 PM


b1469 Catalysis by Ceria and Related Materials

Understanding Ceria-Based Catalytic Materials 83

As deduced from the TD studies reported in Lisowski et al.,367 at 78 K


and PH2 = 0.2 torr, hydrogen dissociation could only be observed on
films grown at 78 K (unsintered films), whereas on those annealed
at T ≥ 320 K it does not occur. TD techniques have also been used to
investigate the adsorption of atomic hydrogen on sintered films at
78 K. In accordance with Lisowski et al.,367 the total amount of hydro-
gen desorbed from the latter experiment is about 50 times larger
than that determined from the former, which clearly indicates that
the surface density of sites able to dissociate H2 molecules on the
surface of an unsintered thin film is rather low.
Figure 2.17 depicts the TD diagram recorded for the desorption
of hydrogen pre-adsorbed in atomic form on a sintered gold film.
There are two peaks, the area of that occurring at the highest tem-
perature (≈ 220 K), peak II, being approximately 30 times larger
than that of peak I. The trace recorded for peak I was interpreted as
due to a first-order process, with a rather low activation energy, Ea =
11 kJ.mol−1, whereas that for peak II was assigned to a second-order

20

15
P·106(torr)

10

0
100 150 200 250
Temperature (K)

Figure 2.17 Thermal desorption of hydrogen pre-adsorbed in atomic form on a


sintered Au thin film. Adapted from Lisowski et al.367 with permission from Elsevier.

b1469_Ch-02.indd 83 4/8/2013 12:26:17 PM


b1469 Catalysis by Ceria and Related Materials

84 J. J. Delgado et al.

desorption process, with Ea = 54 kJ.mol−1. From these results, the


authors concluded that peak I is due to a very weakly adsorbed form
of molecular H2, whereas peak II would account for the recombina-
tion of H chemisorbed species.367 The parameters characterizing
peak II in Fig. 2.17 are similar to those deduced from the TD study
of H2 adsorbed on an unsintered thin film,367 as well as to those
corresponding to the desorption trace for atomic hydrogen pre-
adsorbed on Au{110}(1 × 2).366 We may conclude, accordingly, that
the chemistry of atomic hydrogen interaction with gold single-
crystal and thin-film surfaces is similar. Though the authors367 do not
identify the precise nature of the sites responsible for the dissocia-
tion of H2 on unsintered thin films, this study may reasonably be
considered as one of the first indications of the role played by the
deficiently coordinated surface gold atoms in this process.

2.3.1.2 Preliminary studies on hydrogen adsorption on oxide-supported


gold catalysts
As stressed by Claus,122 by the middle of the 2000s, the number of
studies on the interaction of H2 with oxide-supported gold catalysts
was rather scarce, quantitative information about the hydrogen
adsorbed on the gold phase being almost completely lacking.122 In
these earlier studies on Au/SiO2,369,370 Au/Al2O369 and Au/TiO2370
catalysts, measurable activity for a number of hydrogenation reac-
tions was observed, which might be considered as an indication of
their ability to activate the H2 molecule. However, attempts at meas-
uring, at 298 K or above, the amount of H2 chemisorbed on sup-
ported gold catalysts have often been unsuccessful,369,370 or, as for an
Au/TiO2 sample,370 very small amounts of adsorbed hydrogen,
which could be reversibly desorbed by evacuation at the adsorption
temperature, were measured. These so discouraging early results369,370
probably explain why studies on the chemisorptive properties of
supported gold catalysts were by the middle of the 2000s so scarce.
It is worth outlining, however, that at the time these pioneering
works were published, the procedures for preparing supported gold
catalysts were not well developed, so that they exhibited a very poor

b1469_Ch-02.indd 84 4/8/2013 12:26:18 PM


b1469 Catalysis by Ceria and Related Materials

Understanding Ceria-Based Catalytic Materials 85

gold dispersion. In fact, the catalysts investigated by Sermon et al.369


and Lin et al.370 were prepared by impregnation techniques, which,
with a few exceptions, have been abandoned because they are
considered to lead to very large mean Au particle sizes, typically in
the order of 20–30 nm.
Later, Tamaru and co-workers371 and van Bokhoven and co-
workers260 reported H2 adsorption on Au/Al2O3 samples with a much
better Au dispersion than those of earlier works.369,370 For Tamaru’s
catalyst,371 Au nanoparticles with a mean size of 3 nm were reported,
whereas mean size values ranging from 1.2 nm to 10 nm were deter-
mined for the series of five catalyst samples investigated by van
Bokhoven and co-workers.260 These studies represent a significant
step forward in the understanding of hydrogen adsorption on
oxide-supported gold catalysts. Tamaru and co-workers371 found the
amount of adsorbed H2, determined using a gas circulating system,
at 298 K, to be H/Au = 0.28. In the second case,260 both total and
irreversible adsorption data, at temperatures ranging from 298 K to
523 K, were determined from an analysis of the corresponding volu-
metric isotherms (PH2 = 1–100 kPa). The amount of irreversibly
chemisorbed hydrogen was estimated from the difference between
two consecutive isotherms separated by 2 h evacuation at the corre-
sponding adsorption temperature. For the total adsorption, the
highest H/Au values, 0.66 at 298 K and 0.73 at 373 K, were deter-
mined for the Au/Al2O3 sample with a mean nanoparticle size of
1.4 nm. With the same catalyst, the amounts of irreversibly chem-
isorbed hydrogen were much smaller, H/Au = 0.16 at 298 K and H/
Au = 0.18 at 373 K. In any case, these studies260,371 show that, when
highly dispersed, gold nanoparticles do activate the dissociative
chemisorption of H2.
In situ X-ray absorption spectroscopy (XAS) studies carried out in
parallel with volumetric measurements have revealed that hydrogen
chemisorption induces changes in Au-L3 and Au-L2 X-ray absorption
near-edge structures.260 This observation, in addition to the amount
and relatively strong nature of the chemisorbed hydrogen, gives fur-
ther support to the occurrence of dissociative adsorption by the gold
nanoparticles. Moreover, with a model that assumes the gold

b1469_Ch-02.indd 85 4/8/2013 12:26:18 PM


b1469 Catalysis by Ceria and Related Materials

86 J. J. Delgado et al.

nanoparticle is spherical, Bus et al.260 used the mean size data, as


determined from scanning-transmission electron microscopy (STEM)
and EXAFS, to estimate Aus/Au ratios, i.e. the fraction of the total
gold atoms that are exposed on the surface of nanoparticles. A com-
parison of the Aus/Au data thus estimated with the corresponding
H/Au values, as determined from volumetric studies, revealed that
(H/Au) < (Aus/Au). This observation allowed the authors260 to con-
clude that H2 adsorption could only occur on corner and edged posi-
tions of the alumina-supported Au nanoparticles.
There are two more aspects of the study by van Bokhoven and
co-workers260 deserving some further comment. First, the amount of
hydrogen chemisorbed on the Au/Al2O3 catalysts was found to
increase with the adsorption temperature, which was interpreted as
due to the activated nature of the process. A similar finding had
earlier been reported for hydrogen adsorption on an Au/TiO2 cata-
lyst reduced at 473 K.370 In Bus et al.260 and Lin and Vannice,370 how-
ever, no experimental studies aimed at evaluating the eventual
contribution of hydrogen spilled over the supports are reported.
Moreover, the occurrence of this effect, which had earlier been
observed in some alumina-supported noble metal catalysts,372 and
much more recently on Au/TiO2,256,270,287 had not even been consid-
ered in the analysis of the results commented on above. Second, the
irreversible adsorption data show that evacuation at temperatures
above 373 K is required to ensure the complete elimination of the
pre-adsorbed hydrogen.260 These desorption temperatures are much
higher than those deduced from the TD studies of atomic hydrogen
chemisorbed on single-crystal and thin-film surfaces.366–368,373 A num-
ber of reasons could explain the very different behaviour exhibited
by H forms adsorbed on massive and supported gold phases. A much
stronger H–Au interaction in the latter case might be one of them.
Differences in the kinetics of the desorption process could be a sec-
ond. There is a third reason, which should not be disregarded:
hydrogen desorbed at the highest temperatures could actually come
from the support. In this respect, an additional consideration can be
made. Though not specifically dealing with H2 adsorption, some
comparative studies of CO interaction with massive and supported
gold phases have suggested that the chemical principles governing

b1469_Ch-02.indd 86 4/8/2013 12:26:18 PM


b1469 Catalysis by Ceria and Related Materials

Understanding Ceria-Based Catalytic Materials 87

both processes are similar (López-Haro et al.250 and references


therein). If so, the dramatic difference found for hydrogen desorp-
tion should be considered as a rather unexpected result.
In addition to the chemisorption studies discussed above, cata-
lytic processes like hydrogenation114,260,369–371,374,375 and H2/D2 reac-
tions114,259,260,263,305,376 have been very useful sources of information
about the nature of the hydrogen interaction with supported gold
nanoparticles. In general, these studies are fully consistent with the
ability of gold nanoparticles to activate the dissociation of H2.
Particularly interesting is a recent study of the H2/D2 exchange for
a series of model catalysts consisting of Au nanoparticles with mean
sizes varying in the range 1.3–5.6 nm supported on a rutile TiO2{110}
single-crystal surface.263 The gold loading was the same for the whole
series of Au/TiO2{110} catalysts. Au{111}, Au{311} and TiO2{110}
model surfaces were also investigated as reference systems. The
experiments were performed in a batch reactor under the following
conditions: PH2 = PD2 = 6 torr, and temperatures ranging from 300 K
to 500 K. The exchange reaction could only be observed on the Au/
TiO2{110} samples, the pure metal and oxide single-crystal surfaces
showing no measurable activity. As shown in Fig. 2.18, at 425 K the

Figure 2.18 Plot of rates of HD formation, and TOF data, at 425 K, against mean Au
nanoparticle size for a series of Au/TiO2(110) model catalysts with the same metal
loading. TOF data are referred to Au surface atom at the perimeter of the metal/
support interface. Reproduced from Fujitani et al.263 with permission from Wiley-VCH.

b1469_Ch-02.indd 87 4/8/2013 12:26:18 PM


b1469 Catalysis by Ceria and Related Materials

88 J. J. Delgado et al.

initial rates for the H2/D2 exchange reaction remarkably increase


as the mean Au nanoparticle size decreases, the effect being particu-
larly strong for catalysts with a metal nanoparticle size smaller
than 2 nm.
Interestingly, if the turn over frequency (TOF) data per Au atom
at the perimeter of the metal/support interface are plotted against
the mean nanoparticle size of the corresponding catalyst, Fig. 2.18,
we can see that they are constant, i.e. the TOF values do not depend
on gold dispersion. Moreover, the apparent activation energy for the
exchange reaction was also found to be the same, 36.2–36.6 kJ.mol−1,
for the whole series of catalysts.263 From all these observations, the
authors concluded that the Au atoms at the perimeter of the inter-
face play an essential role as active sites for the dissociation of H2 in
Au/TiO2 systems. This interpretation has received an additional,
very substantial, support from the study of the H2/D2 exchange
carried out for an inverse system consisting of TiOx islands grown on
an Au{111} single-crystal surface.268 In our view, these results strongly
suggest that the oxide support is also involved in the H2 (D2) activa-
tion process. Moreover, these results have led to a better understand-
ing of the mechanism of some selective hydrogenation reactions and
also, very interestingly, of the spillover process on oxide-supported
gold catalysts. In this respect, it is worth noting that a number of
recent studies on Au/TiO2,256,270,287 Au/CeO2,253,270 Au/CeO2–ZnO253
and Au/CeO2–ZrO2241 have unequivocally proved the occurrence of
the hydrogen spillover phenomena at room temperature.
To summarize, recent studies of hydrogen chemisorption, H2/
D2 exchange and hydrogenation reactions on oxide-supported gold
catalysts clearly show that the Au nanoparticles are able to activate
the dissociation of hydrogen molecules. Likewise, they support the
theory that adsorption only occurs on defective Au surface atoms at
corners and edges of the nanoparticles. A second conclusion drawn
from some of these studies is that the dissociation of H2 on supported
gold nanoparticles could be an activated process.260,370
Hydrogen adsorption on both isolated233,258,262,266 and oxide-
supported104,265 Au clusters has also been investigated from a theoreti-
cal point of view. With the help of the increasing power of computer

b1469_Ch-02.indd 88 4/8/2013 12:26:18 PM


b1469 Catalysis by Ceria and Related Materials

Understanding Ceria-Based Catalytic Materials 89

Table 2.3 Some theoretical data for the molecular and dissociative adsorption of
hydrogen on gold single-crystal and cluster surfaces.
Eadsa Eads b
Au Surface (Molecular) (Atomic) Ea Reference
Au{111} −0.9 +41.7 — 262
Au{001} −0.5 +20.2 — 262
Edge atoms of a monoatomic row grown −52.5 −31.2 27.8c 262
on an Au{111} surface
Au25 cluster. Isomer A. Corner −30.1 −25.9 25.0c 262
Au25 cluster. Isomer B. Top −2.1 −9.0 86.0 c
262
Au25 cluster. Isomer B. Corner −3.6 −16.9 9.0c 262
Au7 cluster. Flat structure as in Au{100} — +137.9 188.1 c
258
Au13 cluster. Icosahedron. Corner 0.29 −36.7 30.5d 233
a −1
Notes: Calculated energy for the molecular adsorption of H2 (kJ.molH ). The reference state was the free
2
Au surface and a hydrogen molecule.
b
Calculated energy for the dissociative adsorption of H2 (kJ.molH −1). The reference state was the molecular
2
form of adsorbed H2.
c
Activation energy (kJ.molH2−1) for the dissociation of H2. The reference for the energy of the transition
state was the H2 adsorbed in molecular form.
d
Activation energy (kJ.molH2−1) for the dissociation of H2. The reference for the energy of the transition
state was the free Au cluster and a H2 molecule.

calculations, the amount and quality of information gained has


grown very rapidly in the last few years. There are two major, rather
general, conclusions that may be drawn from these theoretical stud-
ies. First, most of the calculated energies for the dissociative adsorp-
tion of H2 on Au clusters agree on the thermodynamic feasibility of
the process, Table 2.3.
Second, also in agreement with the experimental studies, the
theoretical calculations show that the active sites for the dissociative
adsorption of H2 are localized at corners and edges of the gold
nanoparticles, i.e. they consist of surface atoms with coordination
numbers ≤ 7. As discussed in López-Haro et al.250 and references
therein, a rather similar conclusion has also been drawn for the CO
adsorption on supported gold nanoparticles.
Figure 2.19 shows the energy profiles for the dissociative chem-
isorption of H2 on different gold surfaces. The calculated data

b1469_Ch-02.indd 89 4/8/2013 12:26:18 PM


b1469 Catalysis by Ceria and Related Materials

90 J. J. Delgado et al.

Figure 2.19 Energy profiles for the dissociative chemisorption of H2 on different


Au surfaces. Reproduced from Corma et al.262 with permission from the Royal
Chemical Society.

reported in Table 2.3 were used as the basis for the elaboration of
these profiles.
As can be seen in Fig. 2.19, the specific nanostructure of the
clusters has a remarkable influence not only on the thermodynam-
ics of the H2 dissociation process, but also, very importantly, on its
kinetics. If the difference between the energy calculated for the
transition state and that of the initial state constituted by a free H2
molecule and the corresponding Au cluster is considered, the
results reported in Corma et al.262 clearly show that even for sites
accomplishing the general structural requirements (coordination
number ≤ 7), the activation energy could be significantly different
depending on the morphology of the Au25 cluster, and for the same
morphology, on the specific nanostructure of the surface site. This
introduces some additional, very subtle, elements to the discussion
of chemical interaction of Au clusters with H2. Also, interestingly, in
contrast to the experimental results that suggested the activated
nature of hydrogen dissociation on supported gold nanoparticles,260
the theoretical calculation mentioned above may be interpreted as
implying that, for some specific morphologies, there are sites on
which the dissociation of H2 could be a non-activated process.

b1469_Ch-02.indd 90 4/8/2013 12:26:18 PM


b1469 Catalysis by Ceria and Related Materials

Understanding Ceria-Based Catalytic Materials 91

Hydrogen adsorption on titania-supported gold systems has also


been investigated by periodic density function (DF) calculations.265
The supercells used to model these catalysts consisted of a series of
Au13 clusters with different morphologies seated on the {001} surface
of an anatase (4 × 4) slab. This slab, which contains 144 atoms dis-
tributed in 9 atomic layers, is equivalent to three TiO2 mono-
layers.265 Regarding the morphology of the supported metal phase,
clusters with one (1L), two (2L-A and 2L-B) or three (3L) layers of
Au atoms were considered. Likewise, systems containing stoichio-
metric and oxygen-deficient (one vacancy) TiO2 surfaces were mod-
elled. Some very fine details of the relation existing between the
nanostructure, the electronic properties and chemical behaviour of
these supported Au clusters for hydrogen dissociation were gained
from this study.265 Thus, as noted by Corma and co-workers,265 both
the morphology of the Au cluster and the interaction with the oxide
support induce a considerable heterogeneity in the nanostructural
environment and electronic properties of the cluster atoms, and
consequent to this, on their ability to activate the H2 molecule.
Thus, the net charge on the gold nanoparticles was found to depend
on the redox state of the support, the particles becoming positively
charged when supported on the stoichiometric surface of TiO2 and
negatively charged when seated on an oxygen-deficient surface.
Likewise, the metal/support charge transfer is sensitive to the clus-
ter geometry. Moreover, calculations show that the electronic charge
is unevenly distributed among the different atoms of the clusters.
These are relevant observations because, as proposed by Corma and
co-workers,265,305 the H2 dissociation preferentially occurs on neutral
or slightly charged Au atoms at corner and edge sites. A rather
similar proposal has been made in a very recent FTIR study in which
H2 was chemisorbed on an Au/TiO2 sample previously treated
with CO.256
Particularly remarkable are the theoretical studies of H2 adsorp-
tion on corner sites of Au clusters directly bonded to O atoms of a
titania support.265 Despite their low coordination number and good
accessibility to H2, they are found to be inactive for the dissociation
of H2. This seems to be a general behaviour for Au atoms in

b1469_Ch-02.indd 91 4/8/2013 12:26:18 PM


b1469 Catalysis by Ceria and Related Materials

92 J. J. Delgado et al.

close contact with the support,256,265,305 which allowed Corma and


co-workers265 to conclude that low-coordinated atoms belonging to
the second layer of a gold cluster would be the most active for H2
dissociation. This proposal contrasts with the conclusion drawn
from H2/D2 exchange experiments reported in Fujitani et al.,263 that
the metal atoms at the perimeter of the Au/TiO2 interface, i.e. those
belonging to the first layer of the gold nanoparticles, play a key role
in the process, and therefore in the dissociation of H2. The results
reported in Fujitani et al.,263 however, suggest the direct implication
of the support in the process, a possibility that was not considered in
the theoretical calculations discussed by Corma and co-workers.265

2.3.1.3 Hydrogen adsorption on cerium-based oxide-supported gold


catalysts
Using some rather isolated data, in particular those obtained
from TPR experiments carried out on cerium-based oxide-
supported gold catalysts,39,235,330 valuable information about the
interaction of hydrogen with this family of catalysts was obtained.
However, the number of studies specifically aimed at investigating
hydrogen chemisorption on these catalysts is presently very
scarce.241,253,267,270 Despite this limitation, a number of relevant
conclusions may be drawn. Thus, in Collins et al.241 hydrogen
adsorption by a 3% Au/Ce0.62Zr0.38O2 (Au/CZ) catalyst, with a BET
surface area, 62.8 m2.g−1, was investigated at room temperature
using volumetric adsorption and FTIR techniques. This approach
has an unusual, and certainly very fruitful, combination of tech-
niques. As discussed below, it has rendered interesting data on the
amount and nature of the hydrogen chemisorbed on Au/CZ and
closely related catalysts. Likewise, these techniques have also
provided useful information about the influence of different
catalyst pre-treatments and of CO co-adsorption, on the chemistry
of the H2–(Au/CZ) system.241
Figure 2.20(a) shows the results of the volumetric study. The
experiment used two consecutive isotherms at 308 K, separated by
a 30 min evacuation at 308 K. Prior to running the isotherms, the

b1469_Ch-02.indd 92 4/8/2013 12:26:18 PM


b1469 Catalysis by Ceria and Related Materials

Understanding Ceria-Based Catalytic Materials 93

9.0 700

8.0 1 600
7.0

(µmole H2/g of catalyst)


Apparent (H/Au) ratio
500

Adsorbed amount
6.0
400
5.0

4.0 300

3.0
200
2.0
2 100
1.0

0 50 100 150 200 250 300


PH2 (Torr)
(a)

7
7
2 2 1

(b)

Figure 2.20 (a) Volumetric and (b) FTIR studies of the H2–(Au/CZ) interaction.
(a) Isotherms recorded at 308 K on (1) the as-pre-treated catalyst and (2) the
sample resulting from experiment (1) further evacuated for 30 min at 308 K.
(b) Spectra recorded for (1) pre-treated catalyst and at (2) 3 min, (3) 5 min,
(4) 10 min, (5) 15 min, (6) 20 min and (7) 30 min after admission into the IR cell
of 40 torr of H2 at 298 K. Adapted from Collins et al.241 with permission from the
American Chemical Society.

b1469_Ch-02.indd 93 4/8/2013 12:26:18 PM


b1469 Catalysis by Ceria and Related Materials

94 J. J. Delgado et al.

catalyst sample was heated at 523 K (for 1 h) under 5% O2/He, fol-


lowed by an evacuation at 523 K (for 1 h). From XPS and electron
microscopy studies, this sample consists of Au(0) nanoparticles with
a mean size of 1.8 nm.241,252
On analysing the isotherm (1) in Fig. 2.20(a), a number of
observations may be outlined. At 308 K, the amount of hydrogen
chemisorbed on the Au/CZ catalyst is very large. Under the same
conditions, no measurable adsorption occurs on the bare CZ sup-
port. If expressed as an apparent H/Au ratio, the amount of hydro-
gen chemisorbed on the Au/CZ catalyst, at PH2 = 300 torr, was found
to be H/Au = 8.9. Even at the lowest investigated partial pressure,
PH2 = 5 torr, a H/Au ratio as high as 8.1 was observed. It is obvious
from these quantitative data that at 308 K the spillover is very strong.
Moreover, the estimate in Collins et al.241 suggests that the surface of
the CZ support is saturated by the adsorbed hydrogen. As discussed
at length in Bernal et al.216 and references therein, rather similar
effects were earlier reported for a number of ceria-supported and
ceria–zirconia-supported noble metal catalysts.
The amount of hydrogen chemisorbed for the second of the two
consecutive isotherms is much lower, the measured apparent H/Au
ratio being 0.56, at PH2 = 300 torr. By applying the approach devel-
oped in López-Haro et al.250 to the Au/CZ catalyst investigated in
Collins et al.,241 the total dispersion, D = AuS/AuT = 0.47, i.e. the frac-
tion of the total number of Au atoms exposed at the surface of the
nanoparticles, and the contribution to D of the surface atoms with
coordination (CN ) ≤ 7, D (CN ≤ 7) = AuS(CN ≤ 7)/AuT = 0.26 was
determined by Cíes et al.252 In accordance with both the experimen-
tal and theoretical studies discussed above, the latter dispersion
value, 0.26, could reasonably be used as a measurement of the upper
capability of gold for hydrogen adsorption on the Au/CZ samples
investigated by Collins et al.,241 H/Au = 0.26. If so, the results
obtained for the second isotherm suggest that forms other than the
hydrogen chemisorbed on Au are removed by the 30 min evacuation
treatment at 308 K. These additional forms may consist of spilled-
over hydrogen located in close vicinity of the metal nanoparticles, or
molecular H2 very weakly interacting with the catalyst. Moreover, if

b1469_Ch-02.indd 94 4/8/2013 12:26:18 PM


b1469 Catalysis by Ceria and Related Materials

Understanding Ceria-Based Catalytic Materials 95

earlier H2 chemisorption studies on Au/Al2O3260 are taken into


account, it is likely that not all the hydrogen adsorbed on the gold
nanoparticles is desorbed during the evacuation. In summary,
though valuable, the results reported in Collins et al.241 do not allow
us to make a precise estimate of the contribution of the gold to the
total adsorption data determined from the first isotherm.
The time-resolved FTIR study carried out at 298 K under PH2 =
40 torr, Fig. 2.20(b), fully confirms the conclusions drawn from the
volumetric experiment of Fig. 2.20(a). The activation pre-treatment
applied to the Au/CZ sample was the same in both cases.241 After the
admission of H2 into the infrared (IR) cell, a strong increase in
absorption was noted in the OH stretching region of the spectra. As
deduced from a comparison of the (1) (pre-treated catalyst) and (2)
(after 3 min of contact with H2) spectra, the increase of intensity of
the broad feature centred at 3390 cm−1 characterizing the υ-OH
region of the spectra, is particularly sharp at the beginning of the
experiment, an apparent saturation of the support being reached
after 30 min.241 We may conclude, accordingly, that for Au/CZ, at
room temperature, the spillover is a rather fast process. Numerous
studies of NM(Rh, Pd, Pt)/CeO2(CeO2–ZrO2), Bernal et al.216 and
references therein, and more recently on Au/CeO2253,270 and Au/
TiO2256,287 have confirmed the occurrence of a hydrogen spillover
phenomena at 298 K and even slightly lower temperatures.
There are two more aspects of the FTIR study shown in
Fig. 2.20(b), which merit comment. First, the carbonate region
(1750–1000 cm−1) of the spectra is also modified by the chemisorp-
tion of H2 on Au/CZ, which suggests that the adsorbed carbonate
species present in the sample may also act as a trap for the spilled-
over hydrogen.241 Second, in parallel with the onset of the υ-OH
bands, there is an additional peak at 2133 cm−1. In accordance with
earlier studies,377 this band is assigned to a forbidden electronic
transition 2F5/2 → 2F7/2 in Ce3+.241 The band at 2126 cm−1 reported in
two recent studies for the H2–Au/CeO2 system has been interpreted
in the same way.253,270 Other researchers,267 however, have proposed
an alternative assignment for the band at 2126 cm−1 that they found
in their study of H2 chemisorption on an Au/CeO2 catalyst.

b1469_Ch-02.indd 95 4/8/2013 12:26:18 PM


b1469 Catalysis by Ceria and Related Materials

96 J. J. Delgado et al.

The most commonly accepted interpretation for the IR band at


around 2130 cm−1 implies the transfer of electrons from the spilled-
over hydrogen atoms onto the ceria-containing support, with inher-
ent formation of Ce3+. This effect has been established quantitatively
by a series of magnetic balance studies carried out some time ago on
NM/CeO2 (CeO2–ZrO2) catalysts (Bernal et al.216 and references
therein). This electron transfer from the spilled-over hydrogen to
the titania support has also been firmly evidenced in a number of
recent studies of hydrogen adsorption on Au/TiO2.256,287 This
H-induced modification of the electronic properties of the cerium-
based oxides could be one of the major reasons explaining the
strong downwards shift in their reduction peak often reported in
TPR studies of gold catalysts supported on them.39,235,241,267,330
In Collins et al.,241 some additional FTIR studies of H2 chemisorp-
tion on Au/CZ are reported. They focus on the influence of CO
co-adsorption and of a reducing pre-treatment at 673 K. In both
cases, the authors found a noticeable inhibition of the spillover phe-
nomena. For CO co-adsorption, the conclusion drawn is that H2 and
CO adsorption are competitive processes, that of CO being stronger.
The inhibition effect of the pre-reduction at 673 K is interpreted as
due to the obvious modification induced by the pre-treatment on
the surface properties of the CZ support, the increase in the surface
concentration of oxygen vacancies and the inherent modification of
the electronic properties of the support. No information about the
influence of the support redox state on the chemisorptive proper-
ties of the gold phase against H2 may be deduced from this study.
Theoretical calculations265 and results recently reported for CO
chemisorption for a reduced Au/CZ catalyst,252 however, do not pre-
clude the loss of the H2 adsorption capability of the metal induced
by the reduced support. Also, interestingly, the original H2 chem-
isorptive behaviour of the unreduced Au/CZ catalyst could be
almost completely recovered by re-oxidation at room temperature
of the pre-reduced sample.241
Spectroscopic studies of hydrogen adsorption on Au/CeO2 cata-
lysts recently reported in Tabakova et al.,253 Juárez et al.267 and
Manzoli et al.270 have investigated the nature of the hydrogen species

b1469_Ch-02.indd 96 4/8/2013 12:26:18 PM


b1469 Catalysis by Ceria and Related Materials

Understanding Ceria-Based Catalytic Materials 97

resulting from the process. Juárez et al.267 and Manzoli et al.270


claimed to have detected FTIR bands due to Au–H vibration modes.
In the first case,267 the hydrogen chemisorption experiments were
run at 423 K on a low-loading 0.48% Au/CeO2 sample, which prior
to adsorption had been reduced and further evacuated at 423 K.
These researchers assign to the Au–H mode the band observed at
2126–2128 cm−1. This interpretation is mainly supported by the dis-
appearance of the band on treating the Au/CeO2 catalyst, on which
hydrogen had been pre-adsorbed at 423 K, with O2 at 296 K, or with
CO2 at 423 K. Likewise it was argued by the authors that IR bands at
a similar position had been observed in the IR spectra of matrix
isolated hydride complexes prepared by co-deposition at 3.5 K of
laser-ablated Au atoms with H2.378 It is worth noting, however, that
an FTIR band at an identical position has been observed on
hydrogen-treated Au/CeO2253,270 and Au/CZ241 catalysts, the band
being assigned to a forbidden electronic transition in Ce3+. As previ-
ously discussed in this chapter, the latter interpretation is well estab-
lished in the literature. Despite this, the interpretation above does
not seem to be considered in Juárez et al.267 Moreover, the disappear-
ance of the band after the O2 and CO2 treatments applied by Juárez
et al. might well be interpreted as due to re-oxidation of the Ce3+
species.241,270
By using FTIR spectroscopy, Boccuzzi and co-workers270 also
investigated hydrogen adsorption at room temperature on a 3%
Au/CeO2 catalyst, which had been pre-reduced and evacuated at
423 K. In this work, the behaviour of the Au/CeO2 sample was com-
pared with that exhibited by Au/TiO2 and Au/ZrO2 samples. Some
differences were noted. Electron transfer from the spilled-over
hydrogen to the support was reported to occur for Au/TiO2 and
Au/CeO2. However, for Au/TiO2 this process leads to strongly delo-
calized electrons, as revealed by a very broad absorption ranging
from 3000 cm−1 to 1000 cm−1. After Boccuzzi and co-workers,270 for
Au/CeO2, the electrons trapped by the support are localized in the
form of Ce3+ species, which are identified by the onset of a band at
2126 cm−1. Bands due to H–Au were reported on the three investi-
gated catalysts. The position, however, was found to strongly shift

b1469_Ch-02.indd 97 4/8/2013 12:26:18 PM


b1469 Catalysis by Ceria and Related Materials

98 J. J. Delgado et al.

from 1620 cm−1 for Au/TiO2 and Au/ZrO2, to 1800 cm−1 for Au/
CeO2. The assignment was based on isotopic exchange experiments
with D2 described in the same paper270 and discussed in Juárez
et al.,267 a paper in which the Au–H band was proposed to occur at a
much higher frequency 2126 cm−1. Regarding the origin of the
higher frequency at which the Au–H band is found in Au/CeO2
compared to Au/TiO2 and Au/ZrO2 catalysts, Boccuzzi and co-
workers270 suggested on the basis of theoretical calculations266 that it
is most likely due to the negatively charged nature of the Au nano-
particles dispersed on the partly reduced ceria support.
To summarize, despite the limited number of studies dealing
with the adsorption of H2 on Au/CeO2 and closely related catalysts
which have been published, a number of reasonably well supported
conclusions may be drawn. In good agreement with numerous
theoretical and experimental studies carried out on several
other oxide-supported gold catalysts, particularly on Au/TiO2
samples,256,263,265,270,287,305,308 Au nanoparticles supported on ceria and
ceria-containing mixed oxides are able to activate the dissociation of
H2 at room temperature.241,253,270 Volumetric adsorption and FTIR
studies have revealed that, on Au/CZ, at room temperature, the
chemisorption for Au is accompanied by a fast and very strong
spillover effect leading to the practical saturation of the support,
with the simultaneous onset of an IR band at approximately
2130 cm−1, which is usually interpreted as due to Ce3+ species. This
reduction process implies electron transfer from the spilled-over
atoms of hydrogen to the support. Though not confirmed by experi-
ments on gold catalysts yet, this reduction process is expected to be
partly reversible, i.e. re-oxidation of the support may occur by
simple evacuation of the pre-adsorbed hydrogen under appropriate
conditions. Reversible changes in the redox state of the support
associated with hydrogen adsorption/desorption cycles are known
to occur for a number of cerium-containing oxide-supported noble-
metal catalysts.216
In contrast to CO adsorption on Au/CZ catalysts,250–252 to be dis-
cussed at length in Section 2.3.2 of this chapter, no procedures allow-
ing a quantitative estimate of the metal and support contributions to

b1469_Ch-02.indd 98 4/8/2013 12:26:18 PM


b1469 Catalysis by Ceria and Related Materials

Understanding Ceria-Based Catalytic Materials 99

the total amount of H2 chemisorbed on ceria-containing gold cata-


lysts have been developed as yet. Consequent to this, no reliable data
for the true H/Au value in this family of Au catalysts are presently
available. In good agreement with the wealth of theoretical and
experimental information about H2 adsorption on oxide-supported
gold catalysts discussed above and results already available for CO
adsorption on Au/CZ catalysts, however, it is guessed that, on ceria-
based catalysts, the chemisorption of hydrogen would take place on
defective low-coordination sites of the Au nanoparticles.
The available information, though limited, suggests that the chem-
istry of the H2-(ceria-based Au catalysts) may be significantly modified
by CO co-adsorption and by pre-reduction of the catalysts.241
Recent studies of hydrogen adsorption by Au/CeO2 catalysts267,270
have reported the identification of IR bands attributed to H–Au
modes. Despite the similarity of the pre-treatments applied to the
catalysts, the frequency of this band was found to be very signifi-
cantly shifted from one study, 2126–2128 cm−1,267 to the other 1800
cm−1.270 Such a strong disagreement obliges us to consider these
assignments with some reserve. Additional studies are probably
required to fully confirm the proposals above.

2.3.2 CO adsorption on cerium-based oxide-supported gold catalysts


CO adsorption on gold has been intensively investigated on
single crystals220,227,379–383 and small nanoparticles supported on
both planar model230,232,242,245–247,284,285,298,384–386 and powdered oxi-
des.224–226,234,239,241,249–252,286 From these studies, a first, very general,
conclusion may be drawn: the CO–Au interaction is much weaker
than that observed in noble metals belonging to Groups 8–10.220,387
In agreement with this, it is commonly accepted that CO adsorp-
tion takes place on defective low-coordination surface gold
atoms,220,245,301,387 which has also been confirmed by the energy of
adsorption data determined from theoretical calculations carried
out on using an assumption of different coordination environments
for the gold surface atoms.255,388,389 This behaviour represents a first
important limitation to the use of CO adsorption techniques as a

b1469_Ch-02.indd 99 4/8/2013 12:26:18 PM


b1469 Catalysis by Ceria and Related Materials

100 J. J. Delgado et al.

tool for characterizing the chemical properties of gold catalysts. In


the specific case of catalysts consisting of Au nanoparticles dispersed
on cerium-based oxide supports, there is an additional difficulty that
very much complicates the quantitative analysis of the CO adsorp-
tion data. In effect, it is presently well known that ceria and closely
related mixed oxides may chemisorb large amounts of CO.216
Moreover, as will be discussed later in this section, the amount of CO
adsorbed for this family of supports may be strongly increased by
the presence of a metal phase, and in particular by gold nano-
particles.241,251 There is still an additional source of complication in
the study of CO adsorption for Au/CeO2 and related catalysts: as a
consequence of the reducible nature of the cerium-based oxides,216
SMSI phenomena may eventually occur after applying appropriate
reduction treatments to the catalyst. If so, the chemisorptive behav-
iour of the supported gold nanoparticles may be significantly
disturbed,362 which obviously complicates the analysis of the corre-
sponding quantitative data.
In accordance with the considerations above, it is obvious that
an in-depth analysis of CO adsorption for cerium-based oxide-
supported gold catalysts represents a very challenging issue, which
requires first, the identification of the nature of the main CO forms
occurring on the surface of both the metal phase and the support,
and second, the quantitative establishment of the contribution of
each of these forms to the total amount of adsorbed CO. As will be
discussed below in some detail, these goals have been successfully
achieved in a series of recently published studies in which the chem-
isorption of CO on catalysts consisting of gold nanoparticles
supported on ceria–zirconia mixed oxides was investigated.241,250–252
Regarding the nature of CO adsorbed forms, the curves (a) in
Fig. 2.21 show FTIR spectra recorded at 298 K under PCO = 100 torr,
for the bare support (left) and the corresponding Au/CZ catalyst
(right). As discussed in Collins et al.,241 an analysis of these spectra
allowed the authors to identify three main species. One of them,
which is characterized by an absorption band at 2110 cm−1, was
assigned, in accordance with the literature,239,249,285,344,390 to a linear
form of CO chemisorbed on Au(0) nanoparticles. A second form,

b1469_Ch-02.indd 100 4/8/2013 12:26:18 PM


b1469 Catalysis by Ceria and Related Materials

Understanding Ceria-Based Catalytic Materials 101

b
a
b

Figure 2.21 FTIR study of CO adsorption on CZ support (left) and the corre-
sponding Au/CZ catalyst (right). (a) Spectra recorded for the support and catalyst
samples heated in a flow of 5% O2/He at 523 K (1 h), further evacuated at 523 K
(1 h), and finally treated with 100 torr of CO for 30 min at room temperature. (b)
Spectra recorded after 30 min evacuation at 298 K of samples resulting from (a).
Reproduced from Collins et al.241 with permission from Wiley-VCH.

which could also be identified in the CO stretching region of the


spectra, is characterized by bands occurring above 2143 cm−1. These
bands are attributed to CO adsorbed on cationic sites at the surface
of the CZ support. As deduced from the spectra shown in Fig. 2.21
and marked (b), these two forms are reversible, they are almost
completely removed by a short (30 min) evacuation treatment at
298 K. The very small features at 2125 cm−1 and 2165 cm−1 observed
on the Au/CZ catalyst after this evacuation treatment–Fig. 2.21,
curve (b), right–are interpreted as due to residual cationic forms of
Au not completely reduced during the activation pre-treatment of
the Au/CZ catalyst.
The third form of chemisorbed CO, characterized by bands in
the 1800–1200 cm−1 region of the spectra shown on the right of
Fig. 2.21, is associated with carbonate species resulting from the
CO/CZ interaction. In contrast to the forms mentioned above,
these carbonate species are only very slightly modified by the 30 min
evacuation treatment at 298 K. They are referred to as irreversible
forms of chemisorbed CO.
As discussed in Collins et al.,241 using the results of this FTIR
study the authors proposed a volumetric adsorption routine so that

b1469_Ch-02.indd 101 4/8/2013 12:26:19 PM


b1469 Catalysis by Ceria and Related Materials

102 J. J. Delgado et al.

the contribution of each of the three identified forms to the total


amount of adsorbed CO could be separated and determined quan-
titatively. This routine uses two consecutive isotherms at 308 K, PCO =
0–300 torr, separated by an evacuation treatment of 30 min duration
at 308 K. From the FTIR results shown in Fig. 2.21, the first iso-
therms recorded on the Au/CZ catalyst and the bare support
allowed the authors to determine the total amount of adsorbed CO
on the catalyst and the support, respectively. Likewise, for Au/CZ,
the second isotherm is for the joint contribution of the two reversi-
ble forms, i.e. those due to CO interacting with the Au nanoparticles
and the Ce4+ and Zr+4 sites at the surface of the CZ support. Similarly,
the second of the two consecutive isotherms recorded on the bare
support gives the amount of CO weakly interacting with its cationic
sites. The combination of the results obtained from the parallel
volumetric studies carried out on the catalyst and the bare support
allows us to determine the total amount of CO adsorbed on the cata-
lyst, first isotherm for Au/CZ, the amount of CO weakly adsorbed
on the support cations, second isotherm for CZ, and the amount of
CO chemisorbed on the gold nanoparticles the difference between
the second isotherms recorded for Au/CZ and CZ.
To illustrate this methodology, Fig. 2.22(a) shows the first (1) and
second (2) experimental isotherms for an Au/CZ sample, and the
second isotherm (3) for the corresponding CZ support. From this
study, the difference isotherm (1–2) shown in Fig. 2.22(b) is pro-
posed to account for the CO strongly chemisorbed on the support,
whereas that resulting from the subtraction of isotherms (2) and (3),
(2–3), measures the amount of CO specifically chemisorbed on the
gold nanoparticles.
As outlined in López-Haro et al.,250 the first remarkable conclu-
sion that may be drawn from the shape of the isotherm (2–3) in
Fig. 2.22(b) is that at 308 K the adsorption of CO on the Au nano-
particles reaches a plateau of PCO far below the highest investigated
PCO, i.e. quantitative data at saturation coverage may be determined.
Also, interestingly, the amount of CO strongly chemisorbed on the
CZ support, isotherm (1–2) in Fig. 2.22(b), is much larger than that
determined for CO–Au, isotherm (2–3) in the same figure.

b1469_Ch-02.indd 102 4/8/2013 12:26:19 PM


b1469 Catalysis by Ceria and Related Materials

Understanding Ceria-Based Catalytic Materials 103

140

120 (1)
(µmole CO/g of catalyst)

(1-2)
100
Adsorbed CO

80

60
(2)
40
(3) (2-3)
20

0
0 50 100 150 200 250 300 0 50 100 150 200 250 300
PCO (Torr) PCO (Torr)
(a) (b)

Figure 2.22 Determining the metal and support contributions to the total amount
of CO adsorbed on the catalyst for an AU/CZ sample. (a) First (1) and second
(2) isotherms for Au/CZ, and second isotherm for CZ (3). (b) Difference isotherms
resulting from the subtraction of those in (a): (1–2), and (2–3). Isotherms recorded
at 308 K. Adapted from Cíes et al.251 with permission from Wiley-VCH.

This rather simple experimental routine for separating and


quantifying the contributions of the different forms of CO adsorbed
on conventional powdered Au/CZ and closely related catalysts is a
very powerful tool for investigating their chemical properties.
Moreover, as will be briefly discussed in the next sections of this chap-
ter, the appropriate combination of CO adsorption studies, like
those described above, with the nanostructural information gained
from the application of HRTEM, HAADF-STEM, computer HRTEM
image simulation and computer nanostructural modelling tech-
niques, has allowed the authors to establish rather fine correlations
between the chemical and nanostructural properties in this very chal-
lenging and highly interesting family of gold catalysts.250–252

2.3.2.1 Nanostructural constitution of Au active sites


for CO adsorption on Au/CZ and closely related catalysts
CO adsorption on supported Au nanoparticles is generally acknowl-
edged to occur on low-coordination defective sites; however, this

b1469_Ch-02.indd 103 4/8/2013 12:26:19 PM


b1469 Catalysis by Ceria and Related Materials

104 J. J. Delgado et al.

correlation only received a firm quantitative basis very recently.250 The


approach followed in this study consisted of, first, applying the routine
described above to determine the amount of CO specifically chem-
isorbed, at saturation coverage, on the metal in two catalysts with
significantly different Au nanoparticle size distributions. The catalysts
were 3% Au/Ce0.62Zr0.38O2 (Au/CZ) and 1.5% Au/Ce0.50Tb0.12Zr0.38O2−x
(Au/CTZ).250 In parallel with the chemisorption measurements, the
nanostructural characteristics of the supported metal phase in the
investigated catalysts were carefully analysed.250 In order to establish
the most frequently observed morphology in the supported Au nano-
particles, experimental and computer simulated HRTEM images, like
those shown in Fig. 2.23, were compared. From this study, it was
concluded that the truncated cuboctahedron could be adopted as a
representative morphology for the model Au nanoparticles.
Likewise, by analysing an appropriate number of experimental
STEM-HAADF micrographs, the sizes of approximately 250 Au
nanoparticles were measured, and the corresponding size distribu-
tion determined. The upper part of Fig. 2.24 shows representative

(a) (b) (c)

(d)
8 7
3 4 6 9
0.5nm 0.6nm 1.3 nm
1.9nm 5
2.9nm

Figure 2.23 Representative HRTEM image of a gold nanoparticle supported on a


Ce–ZrTb mixed oxide (a). The corresponding model (b) and image simulation
(c) are also included. (d) A series of gold nanoparticle models of different sizes,
where the coordination numbers are marked. Adapted from López-Haro et al.250
with permission from John Wiley & Sons.

b1469_Ch-02.indd 104 4/8/2013 12:26:19 PM


b1469 Catalysis by Ceria and Related Materials

Understanding Ceria-Based Catalytic Materials 105

30 30

25
D = 0.29 D = 0.69 25
= 1.94 ± 0.01 nm = 1.29 ± 0.03 nm

Frequency (%)
Frequency (%)

20 20

15 1 15

10 10

5 5

0 0
0 1 2 3 4 5 6 10 12 14 16 0 1 2 3 4 5 6 10 12 14 16
Particle Diameter (nm) Particle Diameter (nm)

Figure 2.24 Representative HAADF-STEM images and Au nanoparticle size distri-


butions for Au/CTZ (left) and Au/CZ (right) catalysts. Adapted from López-Haro
et al.250 with permission from John Wiley & Sons.

STEM-HAADF images of Au/CZ and Au/CTZ catalysts. By analysing


these images, the Au nanoparticle size distributions shown in the
lower part of Fig. 2.24 could be determined.250
By using the morphology previously established, the authors250
were able to generate a set of model Au nanoparticles covering the
whole range of sizes observed in the experimental distributions
shown in Fig. 2.24. They analysed the nanostructural characteristics,
i.e. the coordination number, of the Au atoms present at the surface
of each of the model particles, and applied the corresponding
results to the experimental size distributions determined for Au/CZ

b1469_Ch-02.indd 105 4/8/2013 12:26:19 PM


b1469 Catalysis by Ceria and Related Materials

106 J. J. Delgado et al.

and Au/CTZ.250 In this way, some important parameters character-


izing the Au/CZ and Au/CTZ samples could be determined: the
total number of gold atoms in the catalysts (AuT), the total number
of Au atoms at the surface of the nanoparticles (AuS), the number
of surface atoms with coordination number CN = j (AuSj) and the
number of surface atoms localized at the perimeter of the metal/
support interface (AuP). Figure 2.23(d) shows the coordination
number of surface atoms of some model Au nano-particles with
increasing size.
Starting from AuT, AuS, AuSj and Aup, the corresponding disper-
sion parameters, D = AuS/AuT, Dj = AuS,j/AuT, DP = AuP/AuT and Dm-n=
∑m-nDj could be determined. The latter parameter, Dm-n, actually
measures the contribution to the total dispersion, D, of surface
atoms with coordination numbers ranging between two pre-estab-
lished values, j = m and j = n. For the Au/CZ and Au/CTZ catalysts
investigated by López-Haro et al.,250 Au sites were found to show
coordination numbers varying from j = 4 to j = 9. As briefly dis-
cussed below, D4-j values are particularly interesting. Table 2.4 sum-
marizes the results obtained by López-Haro et al.250 using this
methodology. For comparative purposes, in addition to the struc-
tural information for Au surface sites, Table 2.4 includes experi-
mental CO/Au ratios as determined from the volumetric study at
saturation coverage.
As stressed in López-Haro et al.,250 some relevant conclusions
may be drawn from an analysis of Table 2.4. For both catalysts the
experimental CO/Au values are significantly smaller than those

Table 2.4 Contribution of Au surface atoms with CN = j to the metal dispersion in


2.5% Au/Ce0.62Zr0.38O2 (Au/CZ) and 1.5% Au/Ce0.50Tb0.12Zr0.38O2 (Au/CTZ) catalysts
as determined by the methodology developed in López-Haro et al.250
Catalyst CO/Au D4–4 D4–5 D4–6 D4–7 D4–8 D4–9 DP a
Au/CZ 0.49 0.02 0.09 0.26 0.48 0.49 0.69 0.17
Au/CTZ 0.13 0.00 0.02 0.07 0.14 0.16 0.29 0.06
Note: Au atoms at the perimeter of the metal/support interface with CN = j are included in
a

the corresponding Dj values.

b1469_Ch-02.indd 106 4/8/2013 12:26:20 PM


b1469 Catalysis by Ceria and Related Materials

Understanding Ceria-Based Catalytic Materials 107

determined for the total dispersion, D = D4–9, which confirms, as


generally acknowledged,218,220,245,301 that under the investigated
experimental conditions, at saturation coverage only a fraction of
the Au surface atoms is actually involved in the CO adsorption pro-
cess. Interestingly, a comparison of the CO/Au and D4–7 data reveals
that, for both catalysts, the data are in a very good agreement, which
allowed the authors to conclude that the active sites for CO adsorp-
tion actually consist of surface atoms with CN ≤ 7. When the nano-
structural characteristics of these sites were analysed, it was concluded
that they are actually localized at corners, edges and at the perime-
ter of the metal/support interface.
In López-Haro et al.,250 the CO adsorption data determined for
the Au/CZ and Au/CTZ powder catalysts were analysed with refer-
ence to those earlier reported for an Au{110} 1 × 2 reconstructed
model surface.227 The latter study was carried out at much lower
temperatures (100–250K) and CO partial pressures (1 × 10−8 torr to
1 × 10−4torr). As shown by Meier et al.,227 at saturation coverage the
amount of CO chemisorbed on this single-crystal surface is consist-
ent with the exclusive implication of Au sites with CN = 7, i.e. those
located at the edges of the intersection of the small {111} facets that
actually constitute this surface. These edge atoms represent one
fourth of the total number of exposed gold atoms. In accordance
with this comparative analysis, it was concluded by López-Haro
et al.250 that the chemical principles governing the CO/Au interac-
tion in both single-crystal and supported nanoparticle surfaces are
essentially the same, which bridges the material and experimental
gaps existing of the material and experimental conditions of between
both types of studies. Obviously, there are some differences.250 Thus,
in contrast to the ideal single crystal surface, on which no sites with
CN < 7 may be found, on the Au/CZ and Au/CTZ powder catalysts,
Au atoms with CN ≤ 6 may have a significant contribution. In par-
ticular, for the Au/CTZ sample, they make up 24% of the total num-
ber of exposed atoms, and as much as 38% for the Au/CZ catalyst
with the highest metal dispersion.250 A fairly similar conclusion was
reported by Freund and co-workers229 in their study of CO adsorp-
tion on gold nanoparticles of varying size supported on a FeO{111}

b1469_Ch-02.indd 107 4/8/2013 12:26:20 PM


b1469 Catalysis by Ceria and Related Materials

108 J. J. Delgado et al.

film. Recent high-resolution XPS studies of CO adsorption on an


Au/CeO2 catalyst also support this conclusion.245
As discussed in López-Haro et al.,250 some differences in the ther-
modynamics of CO–Au adsorption may also be expected to occur
between the gold single-crystal and oxide-supported nanoparticle
surfaces. According to the literature, the energy of CO adsorption
on gold may depend on variables like the coordination number of
the surface atoms,220,227,229,230 modulations induced on the Au nano-
particles by the nature or redox state of the oxide support232,236,246,252,298
and even, as suggested by some authors,229 on quantum effects asso-
ciated with the size of the nanoparticles. These differences will obvi-
ously determine the specific T–PCO relation at the adsorption
equilibrium.

2.3.2.2 Study of CO strongly chemisorbed on the support in Au/CZ


catalysts: role of the supported metal phase
The experimental routine developed by Collins et al.241 and López-
Haro et al.250 for quantifying the adsorption of CO on Au/CZ cata-
lysts has also been applied to the study of forms strongly chemisorbed
on the support.251 In that study, two Au/CZ samples with a similar
metal loading and the same CZ support were used. The only signifi-
cant difference between the two investigated catalysts was the Au
nano-particle size distribution, and consequent to this, the metal
dispersion, D = AuS/AuT = 0.68 for the sample referred to as Au/
CZ-HD and D = AuS/AuT = 0.49 for the sample referred to as Au/
CZ-MD. Volumetric adsorption data and Au nanoparticle size distri-
butions, as determined by HAADF-STEM, were the experimental
basis for the study carried out by Cíes et al.251 Figure 2.25 summarizes
their results.
Regarding the volumetric adsorption data, the lower part of
Fig. 2.25 shows the (1)–(2) difference isotherms recorded for Au/
CZ-HD, Au/CZ-MD and the bare CZ support. In accordance with
the shape of the three difference isoterms, data at PCO = 100 Torr
were used to estimate the amount of CO strongly chemisorbed
on the CZ support in Au/CZ-HD (220 µmol.g−1), Au/CZ-MD

b1469_Ch-02.indd 108 4/8/2013 12:26:20 PM


b1469 Catalysis by Ceria and Related Materials

Understanding Ceria-Based Catalytic Materials 109

Figure 2.25 Particle size distributions for catalysts Au-MD (upper left) and Au-HD
(upper right). Difference isotherms (lower) for the CO strongly chemisorbed on
CZ for the Au/CZ-MD (ƒ) and Au/CZ-HD (T) catalysts, as well the bare CZ support
(•). Adapted from Cíes et al.251 with permission from John Wiley & Sons.

(97 µmol.g−1) and CZ (6 µmol.g−1).251 As noted by Cíes et al.,251 these


amounts are very much different from one sample to the other,
which is particularly remarkable if we take into account that the CZ
sample is exactly the same in the three cases. Moreover, the amount
of CO adsorbed on the bare support is much smaller than those
determined for the Au/CZ-HD and Au/CZ-MD catalysts, which led

b1469_Ch-02.indd 109 4/8/2013 12:26:20 PM


b1469 Catalysis by Ceria and Related Materials

110 J. J. Delgado et al.

the authors to conclude that the supported gold phase must play a
key role in determining the CO/CZ interaction. Likewise, the
amount of CO adsorbed on Au/CZ-HD is more than double that
determined for Au/CZ-MD. Taking into account that the only sig-
nificant difference between the Au catalysts is their metal disper-
sion, which is higher for the HD sample, it becomes obvious that
metal dispersion, i.e. the Au nanoparticle size distribution, plays an
important role in the process.
Despite the shape of the difference isotherms shown in Fig. 2.25,
which might suggest that a true equilibrium was reached, these
results can only be rationalized by assuming that the irreversible
adsorption of CO on the CZ support is actually controlled by kinetic
factors, and that the presence of the supported Au nanoparticles
makes the first stages of the process much faster.
To understand the influence of metal dispersion on the pro-
cess, a second proposal was made in Cíes et al.251 The growth of the
CO adsorbed phase on the CZ support should follow an annular
model, in which the diffusion of the CO adsorbed on the surface
of the CZ oxide should be the rate-determining step. In accord-
ance with this model, the adsorbed phase would grow around the
Au nanoparticles in the form of annuli, which are assumed to have
the same radius. By using the corresponding Au particle size distri-
bution as an input, Cíes et al.251 were able to make a rough estimate
of this mean radius for both Au/CZ-HD and Au/CZ-MD catalysts.
To simplify their calculations, the surface of the support was
assumed to consist of [111] facets, which have both the highest
thermodynamic stability and the highest surface density of O2−
ions.391 Likewise, in accordance with the stoichiometric require-
ments, one CO molecule was assumed to be adsorbed on two oxide
sites at the surface of the support. Figure 2.26 illustrates the pro-
posed model.
In accordance with the assumptions made, the mean radius
determined for the annuli of the CO phase irreversibly chemisorbed
on the CZ support was found to be 1.7 nm for the Au/CZ-HD sam-
ple and 1.9 nm for the Au/CZ-MD sample. The similarity of the

b1469_Ch-02.indd 110 4/8/2013 12:26:20 PM


b1469 Catalysis by Ceria and Related Materials

Understanding Ceria-Based Catalytic Materials 111

Figure 2.26 Schematic of the annular (blue) model proposed to interpret the
irreversible adsorption of CO on the CZ support in Au/CZ catalysts. Reproduced
from Cíes et al.251 with permission from John Wiley & Sons.

values determined for these two catalysts was considered by Cíes


et al.251 to give some further support to the proposed model.
To confirm the proposed model, some additional assays were
run. The experiment, which was inspired by the H2 and CO co-
adsorption studies reported in Loschen et al.,392 consisted of a FTIR
study of D2 adsorption on an Au/CZ-MD sample, which, prior to its
contact with D2, had been treated with 40 torr of CO at room tem-
perature and further evacuated for 30 min at the same temperature.
As discussed above, CO pre-adsorption should form annuli of car-
bonate species surrounding the Au nanoparticles, whereas the 30
min evacuation would remove the CO adsorbed on the metal with
no significant modification of the carbonate annuli. The Au/CZ-MD
sample pre-treated in this way was finally treated with 40 torr of D2,
at 298 K, and a series of time-resolved FTIR spectra was recorded.
Figure 2.27 shows the results of this study.
As seen in Fig. 2.27, the growth with time of the bands due to
υ-OD stretching modes is much faster on the clean Au/CZ-MD
sample (left) than on the CO-pre-treated one (right). In Cíes
et al.,251 it was estimated that as much as 77% of the total surface of
the support was free from carbonates in the latter sample; there-
fore the authors excluded the lack of available free surface in the

b1469_Ch-02.indd 111 4/8/2013 12:26:20 PM


b1469 Catalysis by Ceria and Related Materials

112 J. J. Delgado et al.

Figure 2.27 Time-resolved FTIR study D2 (PD2 = 40 torr) adsorption at 298 K, on


the clean Au/CZ-MD sample (left), and on the catalyst resulting from a
pre-treatment with 40 torr of CO at 298 K and further evacuated for 30 min at 298 K
(right). The spectra are of (a) the initial sample, after (b) 5 min, (c) 15 min,
(d) 30 min and (e) 60 min in contact with D2. Adapted from Cíes et al.251 with per-
mission from John Wiley & Sons.

support as a likely origin of the different behaviour of the clean


and pre-carbonated catalysts for D2 adsorption. They suggested
instead that the difference was due to perturbations in the kinetics
of the spillover process in the CO-pre-treated catalyst. This pro-
posal is fully consistent with the existence of carbonate annuli
around the Au nanoparticles, which would act as a barrier to the
transfer of the atomic deuterium species from the metal to the
support.
As with CO adsorption on Au nanoparticles briefly discussed
earlier in this chapter, the combination of quantitative data for the CO
irreversibly chemisorbed on the CZ support with nanostructural
information obtained from electron microscopy studies gives a
rather detailed picture of the complex chemistry involved in the
CO adsorption on Au/CZ and closely related catalysts. In addition
to its intrinsic scientific interest, the results reported and discussed
in Cíes et al.251 may certainly inspire some new ideas about the reac-
tion and deactivation mechanisms operating in processes like CO
oxidation, low-temperature water gas shift and PROX, in which
this family of catalysts is known to have very remarkable
properties.

b1469_Ch-02.indd 112 4/8/2013 12:26:20 PM


b1469 Catalysis by Ceria and Related Materials

Understanding Ceria-Based Catalytic Materials 113

2.3.2.3 CO adsorption as a tool for investigating strong metal/support


interaction effects in Au/CZ catalysts
The quantification of the reversible metal deactivation effects
induced on an Au/CZ catalyst by a series of redox pre-treatments,
which consecutively changed the redox state of the support, have
also been investigated using the methodology discussed above.252 In
this investigation, FTIR and volumetric adsorption studies of CO
were combined with nanoparticle size distribution data, as deter-
mined by HAADF-STEM, and redox characterization studies of the
ultimate OSC, as determined by thermo-gravimetric measurements,
and XPS.
In Cíes et al.252 it was reported that a reduction treatment at
473 K strongly disturbs the chemistry of the CO/Au interaction in
an Au/CZ catalyst. Curve (a) on Fig. 2.28(left) shows the υ-CO spec-
trum, recorded at 298 K under PCO = 40 torr for Au/CZ subjected to
an initial treatment under flowing 5% O2/He at 523 K (for 1 h)
followed by 1 h evacuation at 523 K, which was adopted as a stand-
ard activation routine. This was compared to a spectrum recorded
on the catalysts subsequently reduced at 473 K, curve (b), which is
seen to be significantly modified. On the one hand, the IR band is
shifted from 2113 cm−1 in spectrum (a) to 2101 cm−1 in spectrum
(b). On the other hand, the integrated absorption of spectrum
(b) was some 15% smaller than that measured for spectrum (a). As
also shown on the left of Fig. 2.28, the chemical perturbation
induced by the reduction treatment is fully reversible through a
subsequent re-oxidation of sample (b) by applying the initial activa-
tion routine of the catalyst leading to a new spectrum, (c), which is
practically identical to (a).252
From the electron microscopy study reported by Cíes et al.,252 the
modification of the chemical properties mentioned above was not
accompanied by parallel changes either in the total dispersion, D, or
in the D4–7 parameter, which, as already mentioned, would account
for the contribution to D of Au surface sites with CN ≤ 7, i.e. those
which might be expected to be active for CO adsorption. Therefore,
the authors252 concluded that metal sintering/redispersion effects

b1469_Ch-02.indd 113 4/8/2013 12:26:20 PM


b1469 Catalysis by Ceria and Related Materials

114 J. J. Delgado et al.

Figure 2.28 (left) FTIR study of CO adsorption on an AU/CZ sample subjected to


a successive series of redox pre-treatments. (right) Au 4f XP-spectra recorded on
the Au/CZ sample subjected to the same series of pre-treatments. Details of (a), (b)
and (c) pre-treatments are given in the text. Adapted from Cíes et al.252 with permis-
sion from John Wiley & Sons.

were not responsible for the reversible changes that occurred in the
FTIR spectra of the Au/CZ sample after alternating oxidizing and
reducing pre-treatments. They proposed, accordingly, that the most
likely interpretation for the observed effects was an SMSI effect,
similar to that originally reported by Tauster et al.,362,393 i.e. the redox
state of the CZ support reversibly modifies the CO adsorption capa-
bility of the gold nanoparticles supported on it.
In addition to the FTIR study, volumetric measurements pro-
vided quantitative information about the CO chemisorbed on the
metal in the series of catalysts resulting from the three redox pre-
treatments successively applied to the same Au/CZ sample. In
accordance with the methodology developed by López-Haro et al.,250
these quantitative data were determined by subtraction of the sec-
ond isotherms recorded for the catalyst pre-treated as indicated
above, and the corresponding bare support. In this particular case,
because one of the investigated Au/CZ samples had been reduced
at 473 K, the second isotherm corresponding to the bare support
was recorded on an oxide previously reduced at 773 K. As revealed

b1469_Ch-02.indd 114 4/8/2013 12:26:20 PM


b1469 Catalysis by Ceria and Related Materials

Understanding Ceria-Based Catalytic Materials 115

by the ultimate OSC measurements carried out on both Au/CZ and


CZ samples, the redox state of the ceria–zirconia mixed oxide in the
bare support reduced at 773 K and in the Au/CZ catalyst reduced at
473 K, were found to be very close to each other, OSC = 26% for the
gold catalyst reduced at 473 K, and OSC = 29% of the CZ sample
reduced at 773 K.252
By following the procedure discussed above, the amount of CO
specifically chemisorbed on the Au nanoparticles at T = 308 K and
PCO = 40 torr was found to be: 19.4 µmol.g−1 for the initial activated
catalyst, 10.9 µmol.g−1 for the sample reduced at 473 K and
19.3 µmol.g−1 for the catalyst re-oxidized at 523 K.251 These results
give substantial support to the conclusions drawn from the FTIR
study in the sense that a reversible metal deactivation of the Au
nanoparticles does occur on the Au/CZ catalyst successively sub-
jected to oxidizing and reducing pre-treatments. In accordance with
the suggested interpretation of the effect, electron transfer from the
reduced support to the Au nanoparticles is expected. As discussed
in Cíes et al.,252 the Au 4f XP spectra shown on the right of Fig. 2.28
lend further support to the suggested interpretation.
The results presented and discussed in Cíes et al.252 allowed the
authors to draw an additional, very relevant, conclusion. A compari-
son of the integrated absorption data for the υ-CO–Au band
recorded at PCO = 40 torr for the three Au/CZ catalysts, with the cor-
responding volumetric data, also determined at PCO = 40 torr,
revealed that the loss of integrated absorption induced by the reduc-
tion treatment at 473 K was, as already noted, 15%, whereas the
parallel loss of adsorption capability deduced from the volumetric
measurements was as high as 44%. This observation led the authors
to conclude that the absorption coefficient of the υ-CO–Au band is
sensitive to the redox state of the support: the coefficient for the cata-
lyst under the SMSI-like state being larger than that measured for CO
chemisorbed on gold nanoparticles dispersed on a fully oxidized CZ
support. If so, as outlined in Cíes et al.,252 FTIR spectroscopy should
be carefully used in quantitative evaluations of metal deactivation
effects associated with metal/support interactions, which could
modify the electronic structure of the Au nanoparticles.

b1469_Ch-02.indd 115 4/8/2013 12:26:21 PM


b1469 Catalysis by Ceria and Related Materials

116 J. J. Delgado et al.

To summarize, the approach developed in López-Haro et al.250


and Cíes et al.251,252 shows that, despite the usual presumption, CO is
a very interesting probe molecule for characterizing the chemical
properties of gold nanoparticles supported on ceria and ceria-con-
taining mixed oxides. Provided that appropriate routines and com-
binations of experimental techniques are used, the series of studies
discussed in López-Haro et al.250 and Cíes et al.251,252 clearly show that
a rich variety of rather detailed chemical information may be gained
from CO adsorption measurements run under very conventional
experimental conditions. Among several relevant results, the nano-
structural constitution of the surface sites responsible for CO
adsorption by the supported Au nanoparticles, the critical role
played by the metal phase in the strong CO adsorption by the sup-
port and the occurrence of strong fully reversible metal deactivation
effects associated with changes in the redox state of the support,
have been unequivocally established. These studies are a source of
relevant chemical information about Au/CeO2–ZrO2 and closely
related materials; additionally, the discussed results may be fruitfully
used to better understand the reaction and deactivation mecha-
nisms operating in catalytic processes widely investigated for this
family of catalysts.

Acknowledgments
This work was supported by the Spanish Ministry of Science and
Innovation (FEDER/MICINN programme) through the funding of
projects NANOLANCAT (Ref: MAT2008–00889NAN) and IMAGINE
(MAT2009-CSD-00013). Support from Junta de Andalucía (FQM110
and FQM334 groups) is also gratefully acknowledged. X. Chen
acknowledges the Ramón y Cajal program of the Spanish Ministry of
Science and Innovation. Most of the electron microscopy material
included in this chapter was obtained at the Electron Microscopy
Facilities at SCCyT UCA. Data obtained at the University of Cádiz
were completed with different collaborators developed within the
FP6-I3 Project ESTEEM (Contract number RI-I3–026019).

b1469_Ch-02.indd 116 4/8/2013 12:26:21 PM


b1469 Catalysis by Ceria and Related Materials

Understanding Ceria-Based Catalytic Materials 117

References
1. Trovarelli, A., ed., Catalysis by Ceria and Related Materials, Imperial
College Press, London, (2002).
2. Karatzas, X., Jansson, K., Gonzalez, A., Dawody, J., Pettersson, L.J.,
Appl. Catal. B, Environ., 106 (2011) 476–487.
3. Wang, N., Chu, W., Zhang, T., Zhao, X.S., Chem. Eng. J., 170 (2011)
457–463.
4. Sukonket, T., Khan, A., Saha, B., Ibrahim, H., Tantayanon, S., Kumar,
P., Idem, R., Energy Fuels, 25 (2011) 864–877.
5. Halabi, M.H., de Croon, M.H.J.M., van der Schaaf, J., Cobden, P.D.,
Schouten, J.C., Appl. Catal. A, Gen., 389 (2010) 68–79.
6. Ambroise, E., Courson, C., Roger, A.C., Kiennemann, A., Blanchard,
G., Rousseau, S., Carrier, X., Marceau, E., La Fontaine, C., Villain, F.,
Catal. Today, 154 (2010) 133–141.
7. Kambolis, A., Matralis, H., Trovarelli, A., Papadopoulou, C., Appl. Catal.
A, Gen., 377 (2010) 16–26.
8. Bampenrat, A., Meeyoo, V., Kitiyanan, B., Rangsunvigit, P.,
Rirksomboon, T., Appl. Catal. A, Gen., 373 (2010) 154–159.
9. Dantas, S.C., Escritori, J.C., Soares, R.R., Hori, C.E., Chem. Eng. J., 156
(2010) 380–387.
10. Alphonse, P., Ansart, F., J. Colloid Interface Sci., 336 (2009) 658–666.
11. Yuan, Z., Ni, C., Zhang, C., Gao, D., Wang, S., Xie, Y., Okada, A., Catal.
Today, 146 (2009) 124–131.
12. Damyanova, S., Pawelec, B., Arishtirova, K., Martinez Huerta, M.V.,
Fierro, J.L.G., Appl. Catal. B, Environ., 89 (2009) 149–159.
13. Escritori, J.C., Dantas, S.C., Soares, R.R., Hori, C.E., Catal. Commun.,
10 (2009) 1090–1094.
14. Therdthianwong, S., Srisiriwat, N., Therdthianwong, A., Croiset, E.,
Int. J. Hydrogen Energy, 36 (2011) 2877–2886.
15. Vagia, E.C., Lemonidou, A.A., J. Catal., 269 (2010) 388–396.
16. De Rogatis, L., Montini, T., Casula, M.E., Fornasiero, P., J. Alloys
Compd., 451 (2008) 516–520.
17. Birot, A., Epron, F., Descorme, C., Duprez, D., Appl. Catal. B, Environ.,
79 (2008) 17–25.
18. Yamaguchi, D., Tang, L., Wong, L., Burke, N., Trimm, D., Nguyen, K.,
Chiang, K., Int. J. Hydrogen Energy, 36 (2011) 6646–6656.

b1469_Ch-02.indd 117 4/8/2013 12:26:21 PM


b1469 Catalysis by Ceria and Related Materials

118 J. J. Delgado et al.

19. Sim, A., Cant, N.W., Trimm, D.L., Int. J. Hydrogen Energy, 35 (2010)
8953–8961.
20. Ambroise, E., Courson, C., Kiennemann, A., Roger, A.C., Pajot, O.,
Samson, E., Blanchard, G., Top. Catal., 52 (2009) 2101–2107.
21. Laosiripojana, N., Kiatkittipong, W., Assabumrungrat, S., A. I. Ch. E.
J., 57 (2011) 2861–2869.
22. Boullosa-Eiras, S., Zhao, T., Chen, D., Holmen, A., Catal. Today, 171
(2011) 104–115.
23. Sadykov, V.A., Sazonova, N.N., Bobin, A.S., Muzykantov, V.S.,
Gubanova, E.L., Alikina, G.M., Lukashevich, A.I., Rogov, V.A.,
Ermakova, E.N., Sadovskaya, E.M., Mezentseva, N.V., Zevak, E.G.,
Veniaminov, S.A., Muhler, M., Mirodatos, C., Schuurman, Y., van
Veen, A.C., Catal. Today, 169 (2011) 125–137.
24. Silva, F.d., Ruiz, J.A.C., de Souza, K.R., Bueno, J.M.C., Mattos, L.V.,
Noronha, F.B., Hori, C.E., Appl. Catal. A, Gen., 364 (2009) 122–129.
25. Lanza, R., Canu, P., Jaeras, S.G., Appl. Catal. A, Gen., 348 (2008)
221–228.
26. Larrondo, S.A., Kodjaian, A., Fabregas, I., Zimicz, M.G., Lamas, D.G.,
Walsoee de Reca, B.E., Amadeo, N.E., Int. J. Hydrogen Energy, 33
(2008) 3607–3613.
27. Mortola, V.B., Ruiz, J.A.C., Mattos, L.V., Noronha, F.B., Hori, C.E.,
Catal. Today, 133 (2008) 906–912.
28. Reddy, B.M., Thrimurthulu, G., Katta, L., Catal. Lett., 141 (2011)
572–581.
29. Ayastuy, J.L., Gurbani, A., Gonzalez-Marcos, M.P., Gutierrez-Ortiz,
M.A., Appl. Catal. A, Gen., 387 (2010) 119–128.
30. Das, D., Thrimurthulu, G., Lakshmi, K., Reddy, B.M., Int. J. Nanotechnol.,
7 (2010) 1166–1177.
31. Reddy, B.M., Reddy, G.K., Ganesh, I., Ferreira, J.M.F., Catal. Lett., 130
(2009) 227–234.
32. Hosseinpour, N., Khodadadi, A.A., Mortazavi, Y., Bazyari, A., Appl.
Catal. A, Gen., 353 (2009) 271–281.
33. Reddy, B.M., Bharali, P., Seo, Y.H., Prasetyanto, E.A., Park, S.E., Catal.
Lett., 126 (2008) 125–133.
34. Moretti, E., Storaro, L., Talon, A., Lenarda, M., Riello, P., Frattini, R., del
Valle Martinez de Yuso, M., Jimenez-López, A., Rodriguez-Castellon, E.,

b1469_Ch-02.indd 118 4/8/2013 12:26:21 PM


b1469 Catalysis by Ceria and Related Materials

Understanding Ceria-Based Catalytic Materials 119

Ternero, F., Caballero, A., Holgado, J.P., Appl. Catal. B, Environ., 102
(2011) 627–637.
35. Teng, M., Luo, L., Chinese J. Catal., 29 (2008) 1215–1220.
36. Watanabe, K., Miyao, T., Higashiyama, K., Yamashita, H., Watanabe,
M., Catal. Commun., 12 (2011) 976–979.
37. Li, G., Wang, Q., Zhao, B., Shen, M., Zhou, R., J. Hazard. Mater., 186
(2011) 911–920.
38. Wang, Q., Zhao, B., Li, G., Zhou, R., Environ. Sci. Technol., 44 (2010)
3870–3875.
39. Burch, R., Phys. Chem. Chem. Phys., 8 (2006) 5483–5500.
40. Goguet, A., Burch, R., Chen, Y., Hardacre, C., Hu, P., Joyner, R.W.,
Meunier, F., Mun, B.S., Thompsett, D., Tibiletti, D., J. Phys. Chem. C,
111 (2007) 16927–16933.
41. Tibiletti, D., Amieiro-Fonseca, A., Burch, R., Chen, Y., Fisher, J.M.,
Goguet, A., Hardacre, C., Hu, P., Thompsett, D., J. Phys. Chem. B, 109
(2005) 22553–22559.
42. Fu, Q., Saltsburg, H., Flytzani-Stephanopoulos, M., Science, 301 (2003)
935–938.
43. Gayen, A., Boaro, M., de Leitenburg, C., Llorca, J., Trovarelli, A., J.
Catal., 270 (2010) 285–298.
44. Boisen, A., Janssens, T.V.W., Schumacher, N., Chorkendorff, I., Dahl, S.,
J. Mol. Catal. A, Chem., 315 (2010) 163–170.
45. Duarte de Farias, A.M., Nguyen-Thanh, D., Fraga, M.A., Appl. Catal. B,
Environ., 93 (2010) 250–258.
46. de Rivas, B., Lopez-Fonseca, R., Sampedro, C., Gutierez-Ortiz, J.I.,
Appl. Catal. B, Environ., 90 (2009) 545–555.
47. Boaro, M., Vicario, M., Llorca, J., de Leitenburg, C., Dolcetti, G.,
Trovarelli, A., Appl. Catal. B, Environ., 88 (2009) 272–282.
48. Chen, H., Ye, Z., Cui, X., Shi, J., Yan, D., Micropor. Mesopor. Mat., 143
(2011) 368–374.
49. Wang, Q., Li, Z., Zhao, B., Li, G., Zhou, R., J. Mol. Catal. A, Chem., 344
(2011) 132–137.
50. Li, G., Wang, Q., Zhao, B., Zhou, R., Appl. Catal. B, Environ., 105
(2011) 151–162.
51. Zhu, C., Liang, H., Li, S., Hong, Y., Ye, D., Chinese J. Inorg. Chem., 27
(2011) 1093–1100.

b1469_Ch-02.indd 119 4/8/2013 12:26:21 PM


b1469 Catalysis by Ceria and Related Materials

120 J. J. Delgado et al.

52. Wang, Q., Li, G., Zhao, B., Zhou, R., J. Hazard. Mater., 189 (2011)
150–157.
53. Wang, Q., Li, G., Zhao, B., Zhou, R., J. Mol. Catal. A, Chem., 339 (2011)
52–60.
54. Han, Z., Wang, J., Yan, H., Shen, M., Wang, J., Wang, W., Yang, M.,
Catal. Today, 158 (2010) 481–489.
55. Zhao, B., Wang, Q., Li, G., Zhou, R., J. Alloys Compd., 508 (2010)
500–506.
56. Wang, Q., Li, G., Zhao, B., Zhou, R., Appl. Catal. B, Environ., 100
(2010) 516–528.
57. Li, G., Wang, Q., Zhao, B., Zhou, R., J. Mol. Catal. A, Chem., 326 (2010)
69–74.
58. Zhao, B., Li, G., Ge, C., Wang, Q., Zhou, R., Appl. Catal. B, Environ.,
96 (2010) 338–349.
59. Lamacz, A., Krzton, A., Musi, A., Da Costa, P., Catal. Lett., 128 (2009)
40–48.
60. Shen, M., Yang, M., Wang, J., Wen, J., Zhao, M., Wang, W., J. Phys.
Chem. C, 113 (2009) 3212–3221.
61. Zhang, L., Weng, D., Wang, B., Wu, X., Catal. Commun., 11 (2010)
1229–1232.
62. Specchia, S., Finocchio, E., Busca, G., Palmisano, P., Specchia, V.,
J. Catal., 263 (2009) 134–145.
63. Sanchez Escribano, V., del Hoyo Martinez, C., Fernandez López, E.,
Gallardo Amores, J.M., Busca, G., Catal. Commun., 10 (2009) 861–864.
64. Zhang, Y., Deng, J., Zhang, L., Qiu, W., Dai, H., He, H., Catal. Today,
139 (2008) 29–36.
65. de Rivas, B., Lopez-Fonseca, R., Gonzalez-Velasco, J.R., Gutierrez-
Ortiz, J.I., Catal. Commun., 9 (2008) 2018–2021.
66. López-Haro, M., Perez-Omil, J.A., Hernandez-Garrido, J.C., Trasobares,
S., Hungria, A.B., Cíes, J.M., Midgley, P.A., Bayle-Guillemaud, P.,
Martinez-Arias, A., Bernal, S., Delgado, J.J., Calvino, J.J., Chem. Cat.
Chem, 3 (2011) 1015–1027.
67. Wu, X., Liu, S., Weng, D., Lin, F., Catal. Commun., 12 (2011) 345–348.
68. Silva, R.F., DeOliveira, E., de Sousa Filho, P.C., Neri, C.R., Serra, O.A.,
Quimica Nova, 34 (2011) 759–763.
69. Weng, D., Li, J., Wu, X., Si, Z., J. Environ. Sci., China, 23 (2011)
145–150.

b1469_Ch-02.indd 120 4/8/2013 12:26:21 PM


b1469 Catalysis by Ceria and Related Materials

Understanding Ceria-Based Catalytic Materials 121

70. Katta, L., Sudarsanam, P., Thrimurthulu, G., Reddy, B.M., Appl. Catal.
B, Environ., 101 (2010) 101–108.
71. Kolli, T., Kanerva, T., Lappalainen, P., Huuhtanen, M., Vippola, M.,
Kinnunen, T., Kallinen, K., Lepisto, T., Lahtinen, J., Keiski, R.L., Top.
Catal., 52 (2009) 2025–2028.
72. Atribak, I., Bueno-Lopez, A., Garcia-Garcia, A., Top. Catal., 52 (2009)
2088–2091.
73. Reddy, B.M., Rao, K.N., Catal. Commun., 10 (2009) 1350–1353.
74. Atribak, I., Bueno-Lopez, A., Garcia-Garcia, A., J. Mol. Catal. A, Chem.,
300 (2009) 103–110.
75. Atribak, I., Bueno-Lopez, A., Garcia-Garcia, A., J. Catal., 259 (2008)
123–132.
76. Sanchez Escribano, V., Fernandez Lopez, E., Gallardo-Amores, J.M.,
del Hoyo Martinez, C., Pistarino, C., Panizza, M., Resini, C., Busca, G.,
Combust. Flame, 153 (2008) 97–104.
77. Liang, Q., Wu, X., Weng, D., Lu, Z., Catal. Commun., 9 (2008) 202–206.
78. Atribak, I., Bueno-Lopez, A., Garcia-Garcia, A., Catal. Commun., 9
(2008) 250–255.
79. Zhong, J.B., Ma, D., He, X.Y., Li, J.Z., Chen, Y.Q., Appl. Surf. Sci., 256
(2010) 2859–2862.
80. Hu, C., Zhu, Q., Jiang, Z., Powder Technol., 194 (2009) 109–114.
81. Gubanova, E.L., Sadykov, V.A., van Veen, A.C., Mirodatos, C., React.
Kinet. Catal. L., 97 (2009) 349–354.
82. Yang, S., Besson, M., Descorme, C., Appl. Catal. B, Environ., 100 (2010)
282–288.
83. Kim, K.H., Kim, J.R., Ihm, S.K., J. Hazard. Mater., 167 (2009)
1158–1162.
84. Nousir, S., Keav, S., Barbier J., Jr, Bensitel, M., Brahmi, R., Duprez, D.,
Appl. Catal. B, Environ., 84 (2008) 723–731.
85. Yu, Q., Wu, X., Yao, X., Liu, B., Gao, F., Wang, J., Dong, L., Catal.
Commun., 12 (2011) 1311–1317.
86. Atribak, I., Guillen-Hurtado, N., Bueno-Lopez, A., Garcia-Garcia, A.,
Appl. Surf. Sci., 256 (2010) 7706–7712.
87. Liu, L., Yu, Q., Zhu, J., Wan, H., Sun, K., Liu, B., Zhu, H., Gao, F.,
Dong, L., Chen, Y., J. Colloid. Interface Sci., 349 (2010) 246–255.
88. Zhang, R., Teoh, W.Y., Amal, R., Chen, B., Kaliaguine, S., J. Catal., 272
(2010) 210–219.

b1469_Ch-02.indd 121 4/8/2013 12:26:21 PM


b1469 Catalysis by Ceria and Related Materials

122 J. J. Delgado et al.

89. Yu, Q., Liu, L., Dong, L., Li, D., Liu, B., Gao, F., Sun, K., Dong, L.,
Chen, Y., Appl. Catal. B, Environ., 96 (2010) 350–360.
90. Guillen-Hurtado, N., Atribak, I., Bueno-Lopez, A., Garcia-Garcia, A.,
J. Mol. Catal. A, Chem., 323 (2010) 52–58.
91. Atribak, I., Bueno-Lopez, A., Garcia-Garcia, A., Azambre, B., Phys.
Chem. Chem. Phys., 12 (2010) 13770–13779.
92. Adamowska, M., Krzton, A., Najbar, M., Camra, J., Djega-Mariadassou,
G., Da Costa, P., Appl. Catal. B, Environ., 90 (2009) 535–544.
93. Haneda, M., Shinoda, K., Nagane, A., Houshito, O., Takagi, H.,
Nakahara, Y., Hiroe, K., Fujitani, T., Hamada, H., J. Catal., 259 (2008)
223–231.
94. Liu, Y., Murata, K., Inaba, M., Takahara, I., Okabe, K., J. Jpn Petrol.
Inst., 53 (2010) 153–159.
95. Gurbuz, E.I., Kunkes, E.L., Dumesic, J.A., Appl. Catal. B, Environ., 94
(2010) 134–141.
96. Wang, G., Dai, H., Zhang, L., Deng, J., Liu, C., He, H., Au, C.T., Appl.
Catal. A, Gen., 375 (2010) 272–278.
97. Nedyalkova, R., Niznansky, D., Roger, A.C., Catal. Commun., 10 (2009)
1875–1880.
98. Lee, H.J., Park, S., Jung, J.C., Song, I.K., Korean J. Chem. Eng., 28
(2011) 1518–1522.
99. Lee, H.J., Park, S., Song, I.K., Jung, J.C., Catal. Lett., 141 (2011) 531–537.
100. Hu, Y., Liang, T., Zhou, L., Yu, X., Yin, P., Curr. Nanosci., 6 (2010)
666–668.
101. Gaertner, C.A., Serrano-Ruiz, J.C., Braden, D.J., Dumesic, J.A., Ind.
Eng. Chem. Res., 49 (2010) 6027–6033.
102. Postole, G., Chowdhury, B., Karmakar, B., Pinki, K., Banerji, J.,
Auroux, A., J. Catal., 269 (2010) 110–121.
103. Azizi, Y., Petit, C., Pitchon, V., J. Catal., 256 (2008) 338–344.
104. Boronat, M., Corma, A., Langmuir, 26 (2010) 16607–16614.
105. Campo, B., Volpe, M., Ivanova, S., Touroude, R., J. Catal., 242 (2006)
162–171.
106. Campo, B.C., Ivanova, S., Gigola, C., Petit, C., Volpe, M.A., Catal.
Today, 133–135 (2004) 661–666.
107. Cárdenas-Lizana, F., Gómez-Quero, S., Hugon, A., Delannoy, L.,
Louis, C., Keane, M.A., J. Catal., 262 (2009) 235–243.

b1469_Ch-02.indd 122 4/8/2013 12:26:21 PM


b1469 Catalysis by Ceria and Related Materials

Understanding Ceria-Based Catalytic Materials 123

108. Cárdenas-Lizana, F., Gómez-Quero, S., Perret, N., Keane, M.A., Gold
Bull., 42 (2009) 124–132.
109. Chen, C.S., Lin, J.H., Chen, H.W., Appl. Catal. A, Gen., 298 (2006)
161–167.
110. Chen, Y., Qiu, J., Wang, X., Xiu, J., J. Catal., 242 (2006) 227–230.
111. Cárdenas-Lizana, F., de Pedro, Z.M., Gómez-Quero, S., Keane, M.A.,
J. Molec. Catal. A, Chem., 326 (2010) 48–54.
112. Cárdenas-Lizana, F., Gómez-Quero, S., Idriss, H., Keane, M.A.,
J. Catal., 268 (2009) 223–234.
113. Hutchings, G.J., Hall, M.S., Carley, A.F., Landon, P., Solsona, B.E.,
Kiely, C.J., Herzing, A., Makkee, M., Moulijn, J.A., Overweg, A.,
J. Catal., 242 (2006) 71–81.
114. Kartusch, C., van Bokhoven, J.A., Gold Bull., 42 (2009) 343–347.
115. Li, S.-C., Diebold, U., J. Am. Chem. Soc., 132 (2010) 64–66.
116. McEwan, L., Julius, M., Roberts, S., Fletcher, J.C.Q., Gold Bull., 43
(2010) 298–306.
117. Mertens, P., Wahlen, J., Ye, X., Poelman, H., De Vos, D., Catal. Lett.,
118 (2007) 15–21.
118. Milone, C., Trapani, M.C., Galvagno, S., Appl. Catal. A, Gen., 337
(2008) 163–167.
119. Mohr, C., Hofmeister, H., Claus, P., J. Catal., 213 (2003) 86–94.
120. Nikolaev, S.A., Smirnov, V.V., Gold Bull., 42 (2009) 182–189.
121. Okumura, M., Akita, T., Haruta, M., Catal. Today, 74 (2002) 265–269.
122. Claus, P., Appl. Catal. A, Gen., 291 (2005) 222–229.
123. Schimpf, S., Lucas, M., Mohr, C., Rodemerck, U., Bruckner, A.,
Radnik, J., Hofmeister, H., Claus, P., Catal. Today, 72 (2002) 63–78.
124. Segura, Y., Lopez, N., Perez-Ramirez, J., J. Catal., 247 (2007) 383–386.
125. Wang, C.M., Fan, K.N., Liu, Z.P., J. Catal., 266 (2009) 343–350.
126. Haruta, M., Kobayashi, T., Sano, H., Yamada, N., Chem. Lett., 16 (1987)
405–408.
127. Haruta, M., Kobayashi, K., Sano, H., Yamada, N., Chem. Lett., (1987)
405–408.
128. Haruta, M., Kobayashi, K., Tsubota, S., Nakahara, Y., Chem. Express., 3
(1988) 159.
129. Haruta, M., Yamada, N., Kobayashi, T., Iijima, S., J. Catal., 115 (1989)
301–309.

b1469_Ch-02.indd 123 4/8/2013 12:26:21 PM


b1469 Catalysis by Ceria and Related Materials

124 J. J. Delgado et al.

130. Haruta, M., Nature, 437 (2005) 1098–1099.


131. Pennycook, S.J., Varela, M., J. Electron. Microsc., 60 (2011) S213–S223.
132. Takayanagi, K., Kim, S., Lee, S., Oshima, Y., Tanaka, T., Tanishiro, Y.,
Sawada, H., Hosokawa, F., Tomita, T., Kaneyama, T., Kondo, Y., J.
Electron. Microsc., 60 (2011) S239–S244.
133. Gai, P.L., Yoshida, K., Shute, C., Jia, X., Walsh, M., Ward, M.,
Dresselhaus, M.S., Weertman, J.R., Boyes, E.D., Microsc. Res. Tech., 74
(2011) 664–670.
134. Urban, K.W., Science, 321 (2008) 506–510.
135. Pennycook, S.J., Chisholm, M.F., Lupini, A.R., Varela, M., van
Benthem, K., Borisevich, A.Y., Oxley, M.P., Luo, W., Pantelides, S.T.,
Adv. Imag. Elect. Phy., 153 (2008) 327.
136. Tanaka, N., Adv. Imag. Elect. Phy., 153 (2008) 385.
137. Urban, K., Houben, L., Jia, C.L., Lentzen, M., Mi, S.B., Thust, A.,
Tillmann, K., Adv. Imag. Elect. Phy., 153 (2008) 439–480.
138. Varela, M., Lupini, A.R., van Benthem, K., Borisevich, A.Y., Chisholm,
M.F., Shibata, N., Abe, E., Pennycook, S.J., Ann. Rev. Mater. Res., 35
(2005) 539–569.
139. Maigne, A., Twesten, R.D., J. Electron. Microsc., 58 (2009) 99–109.
140. Allen, L.J., Nat. Nanotechnol., 3 (2008) 255–256.
141. Bosman, M., Keast, V.J., Garcia-Muñoz, J.L., D’Alfonso, A.J.,
Findlay, S.D., Allen, L.J., Phys. Rev. Lett., 99 (2007) 086102.
142. Kimoto, K., Asaka, T., Nagai, T., Saito, M., Matsui, Y., Ishizuka, K.,
Nature, 450 (2007) 702–704.
143. Muller, D.A., Fitting Kourkoutis, L., Murfitt, M., Song, J.H., Hwang, H.Y.,
Silcox, J., Dellby, N., Krivanek, O.L., Sci., 319 (2008) 1073–1076.
144. Friedrich, H., de Jongh, P.E., Verkleij, A.J., de Jong, K.P., Chem. Rev.,
109 (2009) 1613–1629.
145. Midgley, P.A., Ward, E.P.W., Hungria, A.B., Thomas, J.M., Chem. Soc.
Rev., 36 (2007) 1477–1494.
146. Thomas, J.M., Midgley, P.A., Chem. Commun., (2004) 1253–1267.
147. Batenburg, K.J., Bals, S., Sijbers, J., Kuebel, C., Midgley, P.A., Hernandez,
J.C., Kaiser, U., Encina, E.R., Coronado, E.A., Van Tendeloo, G.,
Ultramicroscopy, 109 (2009) 730–740.
148. Arslan, I., Walmsley, J.C., Rytter, E., Bergene, E., Midgley, P.A., J. Am.
Chem. Soc., 130 (2008) 5716–5719.

b1469_Ch-02.indd 124 4/8/2013 12:26:21 PM


b1469 Catalysis by Ceria and Related Materials

Understanding Ceria-Based Catalytic Materials 125

149. Ward, E.P.W., Yates, T.J.V., Fernandez, J.J., Vaughan, D.E.W., Midgley,
P.A., J. Phys. Chem. C, 111 (2007) 11501–11505.
150. Midgley, P.A., Thomas, J.M., Laffont, L., Weyland, M., Raja, R.,
Johnson, B.F.G., Khimyak, T., J. Phys. Chem. B, 108 (2004) 4590–4592.
151. Laffont, L., Weyland, M., Raja, R., Thomas, J.M., Midgley, P.A., Elec.
Microsc. Ana., 2003, (2004) 459–462.
152. Midgley, P.A., Weyland, M., Ultramicroscopy, 96 (2003) 413–431.
153. Midgley, P.A., Weyland, M., Thomas, J.M., Gai, P.L., Boyes, E.D.,
Angew. Chem. Int. Ed., 41 (2002) 3804–3807.
154. Weyland, M., Midgley, P.A., Thomas, J.M., J. Phys. Chem. B, 105 (2001)
7882–7886.
155. Midgley, P.A., Weyland, M., Thomas, J.M., Johnson, B.F.G., Chem.
Commun., (2001) 907–908.
156. Liu, J., Chem. Cat. Chem., 3 (2011) 934–948.
157. Thomas, J.M., Midgley, P.A., Chem. Cat. Chem., 2 (2010) 783–798.
158. Krumeich, F., Mueller, E., Wepf, R.A., Nesper, R., J. Phys. Chem. C, 115
(2011) 1080–1083.
159. Hansen, T.W., Wagner, J.B., Dunin-Borkowski, R.E., Mater. Sci. Technol.,
26 (2010) 1338–1344.
160. Nellist, P.D., Kirkland, A.I., Philos. Mag., 90 (2010) 4751–4767.
161. Gai, P.L., Boyes, E.D., Electron Microscopy and Analysis Group
Conference 2009 (Emag 2009), 241 (2010) 012055.
162. Thomas, J.M., Gal, P.L., Adv. Catal., 48 (2004) 171–227.
163. Pijolat, M., Prin, M., Soustelle, M., Touret, O., Nortier, P., J. Chem. Soc.,
Faraday T., 91 (1995) 3941–3948.
164. Cuif, J.P., Blanchard, G., Touret, O., Seigneurin, A., Marczi, M.,
Quemere, E., SAE Technical Paper 970463 (1997).
165. Ozawa, M., Kimura, M., Isogai, A., J. Alloys Compd., 193 (1993) 73–75.
166. Haneda, M., Miki, K., Kakuta, N., Ueno, A., Tani, S., Matsuura, S.,
Sato, M., Nihon Kagaku Kaishi, (1990) 820.
167. Ohata, T., Rare Earths, 17 (1990) 37.
168. Murota, T., Hasegawa, T., Aozasa, S., Matsui, H., Motoyama, M.,
J. Alloys Compd., 193 (1993) 298–299.
169. Rao, G.R., Kašpar, J., Meriani, S., Di Monte, R., Graziani, M., Catal.
Lett., 24 (1994) 107–112.
170. Ozawa, M., Kimura, M., Isogai, A., J. Alloys Compd., 193 (1993) 73–75.

b1469_Ch-02.indd 125 4/8/2013 12:26:21 PM


b1469 Catalysis by Ceria and Related Materials

126 J. J. Delgado et al.

171. Trovarelli, A., Zamar, F., Llorca, J., De Leitenburg, C., Dolcetti, G.,
Kiss, J.T., J. Catal., 169 (1997) 490–502.
172. Rao, G.R., Fornasiero, P., Kašpar, J., Meriani, S., Monte, R.D.,
Graziani, M., Stud. Surf. Sci. Catal., 96 (1995) 631–643.
173. Fornasiero, P., Dimonte, R., Rao, G.R., Kašpar, J., Meriani, S.,
Trovarelli, A., Graziani, M., J. Catal., 151 (1995) 168–177.
174. Fornasiero, P., Fonda, E., Di Monte, R., Vlaic, G., Kašpar, J., Graziani,
M., J. Catal., 187 (1999) 177–185.
175. Vlaic, G., Di Monte, R., Fornasiero, P., Fonda, E., Kašpar, J., Graziani,
M., J. Catal., 182 (1999) 378–389.
176. Boaro, M., De Leitenburg, C., Dolcetti, G., Trovarelli, A., J. Catal., 193
(2000) 338–347.
177. Hickey, N., Fornasiero, P., Kašpar, J., Graziani, M., Blanco, G., Bernal,
S., Chem. Commun., (2000) 357–358.
178. Hickey, N., Fornasiero, P., Kašpar, J., Gatica, J.M., Bernal, S., J. Catal.,
200 (2001) 181–193.
179. Baker, R.T., Bernal, S., Blanco, G., Cordon, A.M., Pintado, J.M.,
Rodriguez-Izquierdo, J.M., Fally, F., Perrichon, V., Chem. Commun.,
(1999) 149–150.
180. Fornasiero, P., Balducci, G., Kašpar, J., Meriani, S., Di Monte, R.,
Graziani, M., Catal. Today, 29 (1996) 47–52.
181. Fornasiero, P., Balducci, G., Di Monte, R., Kašpar, J., Sergo, V.,
Gubitosa, G., Ferrero, A., Graziani, M., J. Catal., 164 (1996) 173–183.
182. Perrichon, V., Laachir, A., Bergeret, G., Fréty, R., Tournayan, L.,
Touret, O., J. Chem. Soc., Faraday T., 90 (1994) 773–781.
183. Laachir, A., Perrichon, V., Badri, A., Lamotte, J., Catherine, E.,
Lavalley, J.C., El Fallah, J., Hilaire, L., Le Normand, F., Quéméré, E.,
Sauvion, G.N., Touret, O., J. Chem. Soc., Faraday T., 87 (1991)
1601–1609.
184. Izu, N., Omata, T., Otsuka-Yao-Matsuo, S., J. Alloys Compd., 270 (1998)
107–114.
185. Fornasiero, P., Montini, T., Graziani, M., Kašpar, J., Hungría, A.B.,
Martínez-Arias, A., Conesa, J.C., Phys. Chem. Chem. Phys., 4 (2002)
149–159.
186. Montini, T., Bañares, M.A., Hickey, N., Di Monte, R., Fornasiero, P.,
Kašpar, J., Graziani, M., Phys. Chem. Chem. Phys., 6 (2004) 1–3.

b1469_Ch-02.indd 126 4/8/2013 12:26:21 PM


b1469 Catalysis by Ceria and Related Materials

Understanding Ceria-Based Catalytic Materials 127

187. Montini, T., Hickey, N., Fornasiero, P., Graziani, M., Bañares, M.A.,
Martínez-Huerta, M.V., Alessandri, I., Depero, L.E., Chem. Mater., 17
(2005) 1157–1166.
188. Yashima, M., Arashi, H., Kakihana, M., Yoshimura, M., J. Am. Ceram.
Soc., 77 (1994) 1067–1071.
189. Yashima, M., Sasaki, S., Yamaguchi, Y., Kakihana, M., Yoshimura, M.,
Mori, T., Appl. Phys. Lett., 72 (1998) 182–184.
190. Kishimoto, H., Omata, T., Otsuka-Yao-Matsuo, S., Ueda, K., Hosono,
H., Kawazoe, H., J. Alloys Compd., 312 (2000) 94–103.
191. Otsuka-Yao-Matsuo, S., Omata, T., Izu, N., Kishimoto, H., J. Solid State
Chem., 138 (1998) 47–54.
192. Perez Omil, J.A., Bernal, S., Calvino, J.J., Hernández, J.C., Mira, C.,
Rodríguez-Luque, M.P., Erni, R., Browning, N.D., Chem. Mater., 17
(2005) 4282–4285.
193. Sasaki, T., Ukyo, Y., Kuroda, K., Arai, S., Saka, H., J. Electron. Microsc.,
52 (2003) 309–312.
194. Sasaki, T., Ukyo, Y., Suda, A., Sugiura, M., Kuroda, K., Arai, S., Saka,
H., J. Ceram. Soc. Japan, 111 (2003) 382–385.
195. Torng, S., Miyazawa, K., Suzuki, K., Sakuma, T., Philos. Mag. A, 70 (1994)
505.
196. Masui, T., Ozaki, T., Adachi, G., Kang, Z., Eyring, L., Chem. Lett.,
(2000) 840–841.
197. Hernández, J.C., Hungría, A.B., Pérez-Omil, J.A., Trasobares, S.,
Bernal, S., Midgley, P.A., Alavi, A., Calvino, J.J., J. Phys. Chem. C, 111
(2007) 9001–9004.
198. Pennycook, S.J., Jesson, D.E., Ultramicroscopy, 37 (1991), 14.
199. Trasobares, S., López-Haro, M., Kociak, M., March, K., de la Peña, F.,
Perez-Omil, J.A., Calvino, J.J., Lugg, N.R., D’Alfonso, A.J., Allen, L.J.,
Colliex, C., Angew. Chem. Int. Ed., 50 (2011) 868–872.
200. Izu, N., Kishimoto, H., Omata, T., Yao, T., Otsuka-Yao-Matsuo, S., Sci.
Technol. Adv. Mat., 2 (2001) 443–448.
201. Izu, N., Kishimoto, H., Omata, T., Ono, K., Otsuka-Yao-Matsuo, S., Sci.
Technol. Adv. Mat., 2 (2001) 397–404.
202. Colon, G., Pijolat, M., Valdivieso, F., Vidal, H., Kašpar, J., Finocchio, E.,
Daturi, M., Binet, C., Lavalley, J.C., Baker, R.T., Bernal, S., J. Chem. Soc.
Faraday T., 94 (1998) 3717–3726.

b1469_Ch-02.indd 127 4/8/2013 12:26:21 PM


b1469 Catalysis by Ceria and Related Materials

128 J. J. Delgado et al.

203. Colon, G., Valdivieso, F., Pijolat, M., Baker, R.T., Calvino, J.J., Bernal,
S., Catal. Today, 50 (1999) 271–284.
204. Shah, P.R., Kim, T., Zhou, G., Fornasiero, P., Gorte, R.J., Chem. Mater.,
18 (2006) 5363–5369.
205. Zhou, G., Shah, P.R., Kim, T., Fornasiero, P., Gorte, R.J., Catal. Today,
123 (2007) 86–93.
206. Wang, H.-F., Guo, Y.-L., Lu, G.-Z., Hu, P., Angew. Chem. Int. Ed., 48
(2009) 8289–8292.
207. Yeste, M.P., Hernández, J.C., Bernal, S., Blanco, G., Calvino, J.J., Pérez-
Omil, J.A., Pintado, J.M., Chem. Mater., 18 (2006) 2750–2757.
208. Yeste, M.P., Hernandez, J.C., Trasobares, S., Bernal, S., Blanco, G.,
Calvino, J.J., Perez-Omil, J.A., Pintado, J.M., Chem. Mater., 20 (2008)
5107–5113.
209. Bernal, S., Calving, J.J., Cifredo, G.A., Rodríguez-Izquierdo, J.M.,
Perrichon, V., Laachir, A., J. Catal., 137 (1992) 1–11.
210. Bernal, S., Calvino, J.J., Cifredo, G.A., Rodríguez-Izquierdo, J.M.,
J. Phys. Chem., 99 (1995) 11794–11796.
211. Norman, A., Perrichon, V., Bensaddik, A., Lemaux, S., Bitter, H.,
Koningsberger, D., Top. Catal., 16–17 (2001) 363–368.
212. Bernal, S., Blanco, G., Calvino, J.J., Hernández, J.C., Pérez-Omil, J.A.,
Pintado, J.M., Yeste, M.P., J. Alloys Compd., 451 (2008) 521–525.
213. Yeste, M.P., Hernandez, J.C., Bernal, S., Blanco, G., Calvino, J.J., Pérez-
Omil, J.A., Pintado, J.M., Catal. Today, 141 (2009) 409–414.
214. Mondelli, C., Santo, V.D., Trovarelli, A., Boaro, M., Fusi, A., Psaro, R.,
Recchia, S., Catal. Today, 113 (2006) 81–86.
215. Boaro, M., Giordano, F., Recchia, S., Santo, V.D., Giona, M., Trovarelli, A.,
Appl. Catal. B, Environ., 52 (2004) 225–237.
216. Bernal, S., Calvino, J.J., Gatica, J.M., López-Cartes, C., Pintado, J.M.,
in Catalysis by Ceria and Related Materials, ed. A. Trovarelli, Imperial
College Press, London, (2002), pp. 85–168.
217. Hutchings, G.J., Dalton T., (2008) 5523–5536.
218. Bond, G.C., Louis, C., Thompson, D.T., Catalysis by Gold, Imperial
College Press, London, (2006).
219. Bond, G.C., Thompson, D.T., Catal. Rev. Sci. Eng., 41 (1999) 319–388.
220. Meyer, R., Lemire, C., Shaikhutdinov, S.K., Freund, H.J., Gold Bull., 37
(2004) 72–124.

b1469_Ch-02.indd 128 4/8/2013 12:26:21 PM


b1469 Catalysis by Ceria and Related Materials

Understanding Ceria-Based Catalytic Materials 129

221. Carabineiro, S.A.C., Nieuwenhuys, B.E., Gold Bull., 42 (2009)


288–301.
222. Carabineiro, S.A.C., Nieuwenhuys, B.E., Gold Bull., 43 (2010) 262–266.
223. Delgado, J.J., Cíes, J.M., López-Haro, M., del Río, E., Calvino, J.J.,
Bernal, S., Chem. Lett., 40 (2011) 1210–1216.
224. Boccuzzi, F., Chiorino, A., Manzoli, M., Andreeva, D., Tabakova, T.,
J. Catal., 188 (1999) 176–185.
225. Boccuzzi, F., Chiorino, A., Manzoli, M., Lu, P., Akita, T., Ichikawa, S.,
Haruta, M., J. Catal., 202 (2001) 256–267.
226. Manzoli, M., Chiorino, A., Boccuzzi, F., Surf. Sci., 532–535 (2003)
377–382.
227. Meier, D.C., Bukhtiyarov, V., Goodman, D.W., J. Phys. Chem. B, 107
(2003) 12668–12671.
228. Shaikhutdinov, S.K., Meyer, R., Naschitzki, M., Baumer, M., Freund,
H.J., Catal. Lett., 86 (2003) 211–219.
229. Lemire, C., Meyer, R., Shaikhutdinov, S., Freund, H.J., Surf. Sci., 552
(2004) 27–34.
230. Meier, D.C., Goodman, D.W., J. Am. Chem. Soc., 126 (2004) 1892–1899.
231. Derrouiche, S., Gravejat, P., Bianchi, D., J. Am. Chem. Soc., 126 (2004)
13010–13015.
232. Yoon, B., Haakkinen, H., Landman, U., Wörz, A.S., Antonietti, J.M.,
Abbet, S., Juadai, K., Heiz, U., Science, 307 (2005) 403–407.
233. Okumura, M., Kitagawa, Y., Haruta, M., Yamaguchi, K., Appl. Catal. A,
Gen., 291 (2005) 37–44.
234. Menegazzo, F., Manzoli, M., Chiorino, A., Boccuzzi, F., Tabakova, T.,
Signoretto, M., Pinna, F., Pernicone, N., J. Catal., 237 (2006) 431–434.
235. Manzoli, M., Boccuzzi, F., Chiorino, A., Vindigni, F., Deng, W.,
Flytzani-Stephanopoulos, M., J. Catal., 245 (2007) 308–315.
236. Sterrer, M., Yulikov, M., Fischbach, E., Heyde, M., Rust, H.-P.,
Pacchioni, G., Risse, T., Freund, H.-J., Angew. Chem. Int. Ed., 45 (2006)
2630–2632.
237. Sterrer, M., Yulikov, M., Risse, T., Freund, H.-J., Carrasco, J., Illas, F.,
Di Valentin, C., Giordano, L., Pacchioni, G., Angew. Chem. Int. Ed., 45
(2006) 2633–2635.
238. Taylor, B., Lauterbach, J., Blau, G.E., Delgass, W.N., J. Catal., 242
(2006) 142–152.

b1469_Ch-02.indd 129 4/8/2013 12:26:21 PM


b1469 Catalysis by Ceria and Related Materials

130 J. J. Delgado et al.

239. Concepcion, P., Carrettin, S., Corma, A., Appl. Catal. A, Gen., 307
(2006) 42–45.
240. Tabakova, T., Boccuzzi, F., Manzoli, M., Sobczak, J.W., Idakiev, V.,
Andreeva, D., Appl. Catal. A, Gen., 298 (2006) 127–143.
241. Collins, S.E., Cíes, J.M., del Río, E., López-Haro, M., Trasobares, S.,
Calvino, J.J., Pintado, J.M., Bernal, S., J. Phys. Chem. C, 111 (2007)
14371–14379.
242. Diemant, T., Hartmann, H., Bansmann, J., Behm, R.J., J. Catal., 252
(2007) 171–177.
243. Chen, M., Goodman, D.W., Chem. Soc. Rev., 37 (2008) 1860–1870.
244. Avgouropoulos, G., Manzoli, M., Boccuzzi, F., Tabakova, T.,
Papavasiliou, J., Ioannides, T., Idakiev, V., J. Catal., 256 (2008) 237–247.
245. Weststrate, C.J., Resta, A., Westerström, R., Lundgren, E., Mikkelsen, A.,
Andersen, J.N., J. Phys. Chem. C, 112 (2008) 6900–6906.
246. Weststrate, C.J., Westerström, R., Lundgren, E., Mikkelsen, A.,
Andersen, J.N., Resta, A., J. Phys. Chem. C, 113 (2009) 724–728.
247. Baron, M., Bondarchuk, O., Stacchiola, D., Shaikhutdinov, S., Freund,
H.J., J. Phys. Chem. C, 113 (2009) 6042–6049.
248. Menegazzo, F., Pinna, F., Signoretto, M., Trevisan, V., Boccuzzi, F.,
Chiorino, A., Manzoli, M., Appl. Catal. A, Gen., 356 (2009) 31–35.
249. Vindigni, F., Manzoli, M., Chiorino, A., Boccuzzi, F., Gold Bull., 42
(2009) 106–112.
250. López-Haro, M., Delgado, J.J., Cíes, J.M., del Río, E., Bernal, S.,
Burch, R., Cauqui, M.A., Trasobares, S., Pérez-Omil, J.A., Bayle-
Guillemaud, P., Calvino, J.J., Angew. Chem. Int. Ed., 49 (2010) 1981–1985.
251. Cíes, J.M., Delgado, J.J., López-Haro, M., Pilasombat, R., Trasobares, S.,
Bernal, S., Calvino, J.J., Chem. Eur. J., 16 (2010) 9536–9543.
252. Cíes, J.M., del Río, E., López-Haro, M., Delgado, J.J., Blanco, G.,
Collins, S.E., Calvino, J.J., Bernal, S., Angew. Chem. Int. Ed., 49 (2010)
9744–9748.
253. Tabakova, T., Manzoli, M., Vindigni, F., Idakiev, V., Boccuzzi, F., J. Phys.
Chem. A, 114 (2010) 3909–3915.
254. Brown, M.A., Fujimori, Y., Ringle, F., Shao, X., Stavale, F., Nilius, N.,
Sterrer, M., Freund, H.-J., J. Am. Chem. Soc., 133 (2011) 10668–10676.
255. Sun, K., Kohyama, M., Tanaka, S., Takeda, S., J. Comput. Chem., 32
(2011) 3276–3282.

b1469_Ch-02.indd 130 4/8/2013 12:26:21 PM


b1469 Catalysis by Ceria and Related Materials

Understanding Ceria-Based Catalytic Materials 131

256. Panayotov, D.A., Burrows, S.P., Yates Jr., J.T., Morris, J.R., J. Phys. Chem.
C, 115 (2011) 22400–22408.
257. Hammer, B., Norskov, J.K., Nature, 376 (1995) 238–240.
258. Stromsnes, H., Jusuf, S., Schimmelpfennig, B., Wahlgren, U., Gropen, O.,
J. Mol. Struct., 567–568 (2001) 137–143.
259. Gluhoi, A.C., Vreeburg, H.S., Bakker, J.W., Nieuwenhuys, B.E., Appl.
Catal. A, Gen., 291 (2005) 145–150.
260. Bus, E., Miller, J.T., van Bokhoven, J.A., J. Phys. Chem. B, 109 (2005)
14581–14587.
261. Bus, E., van Bokhoven, J.A., Phys. Chem. Chem. Phys., 9 (2007)
2894–2902.
262. Corma, A., Boronat, M., González, S., Illas, F., Chem. Commun., (2007)
3371–3373.
263. Fujitani, T., Nakamura, I., Akita, T., Okumura, M., Haruta, M., Angew.
Chem. Int. Ed., 48 (2009) 9515–9518.
264. Guzmán, C., Del Angel, G., Gómez, R., Galindo, F., Zanella, R.,
Torres, G., Angeles-Chavez, C., Fierro, J.L.G., J. Nano. Res., 5 (2009)
13–23.
265. Boronat, M., Illas, F., Corma, A., J. Phys. Chem. A, 113 (2009)
3750–3757.
266. Zhao, S., Ren, Y., Ren, Y., Wang, J., Yin, W., J. Phys. Chem. A, 114 (2010)
4917–4923.
267. Juárez, R., Parker, S.F., Concepción, P., Corma, A., García, H., Chem.
Sci., 1 (2010) 731–738.
268. Nakamura, I., Mantoku, H., Furukawa, T., Fujitani, T., J. Phys. Chem. C,
115 (2011) 16074–16080.
269. Joshi, A.M., Delgass, W.N., Thomson, K.T., Top. Catal., 44 (2011)
27–39.
270. Manzoli, M., Chiorino, A., Vindigni, F., Boccuzzi, F., Catal. Today, 181
(2012) 62–67.
271. Oh, H.S., Yang, J.H., Costello, C.K., Wang, Y.M., Bare, S.R., Kung, H.H.,
Kung, M.C., J. Catal., 210 (2002) 375–386.
272. Guzmán, J., Carrettin, S., Corma, A., J. Am. Chem. Soc., 127 (2005)
3286–3287.
273. Guzmán, J., Carrettin, S., Fierro-González, J.C., Hao, Y., Gates, B.C.,
Corma, A., Angew. Chem. Int. Ed., 44 (2005) 4778–4781.

b1469_Ch-02.indd 131 4/8/2013 12:26:21 PM


b1469 Catalysis by Ceria and Related Materials

132 J. J. Delgado et al.

274. Kotobuki, M., Leppelt, R., Hansgen, D.A., Widmann, D., Behm, R.J.,
J. Catal., 264 (2009) 67–76.
275. Tost, A., Widmann, D., Behm, R.J., J. Catal., 266 (2009) 299–307.
276. Widmann, D., Liu, Y., Schuth, F., Behm, R.J., J. Catal., 276 (2010)
292–305.
277. Roldan, A., Ricart, J.M., Illas, F., Pacchioni, G., Phys. Chem. Chem. Phys.,
12 (2010) 10723–10729.
278. Roldán, A., Ricart, J.M., Illas, F., Theor. Chem. Acc., 128 (2011)
675–681.
279. Boronat, M., Corma, A., Dalton T., 39 (2010) 8538–8546.
280. Wang, L.C., Jin, H.J., Widmann, D., Weissmüller, J., Behm, R.J.,
J. Catal., 278 (2011) 219–227.
281. Widmann, D., Behm, R.J., Angew. Chem. Int. Ed., 50 (2011)
10241–10245.
282. Fu, Q., Wagner, T., Surf. Sci. Reports, 62 (2007) 431–498.
283. Senanayake, S.D., Stacchiola, D., Liu, P., Mullins, C.B., Hrebk, J.,
Rodriguez, J.A., J. Phys. Chem. C, 113 (2009) 19536–19544.
284. Chen, M., Goodman, D.W., Acc. Chem. Res., 39 (2006) 739–746.
285. Risse, T., Shaikhutdinov, S., Nilius, N., Sterrer, M., Freund, H.-J., Acc.
Chem. Res., 41 (2008) 949–956.
286. Iizuka, Y., Fujiki, H., Ymauchi, N., Chijiiwa, T., Arai, S., Tsubota, S.,
Haruta, M., Catal. Today, 36 (1997) 115–123.
287. Panayotov, D.A., Yates, J.T., J. Phys. Chem. C, 111 (2007) 2959–2964.
288. Chiorino, A., Manzoli, M., Menegazzo, F., Signoretto, M., Vindigni, F.,
Pinna, F., Boccuzzi, F., J. Catal., 262 (2009) 169–176.
289. Roze, E., Quinet, E., Caps, V., Bianchi, D., J. Phys. Chem. C, 113 (2009)
8194–8200.
290. Juárez, R., Parker, S.F., Concepción, P., Corma, A., García, H., Chem.
Sci., 1 (2010) 731–738.
291. Green, I.X., Tang, W., Neurock, M., Yates, J.T., Science, 333 (2011)
736–739.
292. Zhou, Z., Flytzani-Stephanopoulos, M., Saltsburg, H., J. Catal., 280
(2011) 255–263.
293. Phala, N.S., Klatt, G., Steen, E.v., Chem. Phys. Lett., 395 (2004) 33–37.
294. Varganov, S.A., Olson, R.M., Gordon, M.S., Mills, G., Metiu, H.,
J. Chem. Phys., 120 (2004) 5169–5175.

b1469_Ch-02.indd 132 4/8/2013 12:26:21 PM


b1469 Catalysis by Ceria and Related Materials

Understanding Ceria-Based Catalytic Materials 133

295. Loffreda, D., Piccolo, L., Sautet, P., Phys. Rev. B, 71 (2005)
113414-1–4.
296. Barnard, A.S., Lin, X.M., Curtiss, L.A., J. Phys. Chem. B, 109 (2005)
24465–24472.
297. Fielicke, A., von Helden, G., Meijer, G., Pedersen, D.B., Simard, B.,
Rayner, D.M., J. Am. Chem. Soc., 127 (2005) 8416–8423.
298. Wörz, A.S., Heiz, U., Cinquini, F., Pacchioni, G., J. Phys. Chem. B, 109
(2005) 18418–18426.
299. Rashkeev, S.N., Lupini, A.R., Overbury, S.H., Pennycook, S.J.,
Pantelides, S.T., Phys. Rev. B, 76 (2007) 035438-1–8.
300. Hvolbaek, B., Janssens, T.V.W., Clausen, B.S., Falsig, H., Christensen,
C., Norskov, J.K., Nanotoday, 2 (2007) 14–18.
301. Janssens, T.V.W., Clausen, B.S., Hvolbaek, B., Falsig, H., Christensen,
C., Bligaard, T., Norskov, J.K., Top. Catal., 44 (2007) 15–26.
302. Huang, W., Ji, M., Dong, C.-D., Gu, X., Wang, L.-M., Gong, X.G.,
Wang, L.-S., ACS Nano, 2 (2008) 897–904.
303. Ojifinni, R.A., Froemming, N.S., Gong, J., Pan, M., Kim, T.S., White,
J.M., Henkelman, G., Mullins, C.B., J. Am. Chem. Soc., 130 (2008)
6801–6812.
304. Coquet, R., Howard, K.L., Willock, D.J., Chem. Soc. Rev., 37 (2008)
2046–2076.
305. Boronat, M., Concepción, P., Corma, A., J. Phys. Chem. C, 113 (2009)
16772–16784.
306. Zhu, W.J., Zhang, J., Gong, X.Q., Lu, G., Catal. Today, 165 (2011)
19–24.
307. Brodersen, S.H., Gronbjerg, U., Hvolbaek, B., Schiotz, J., J. Catal., 284
(2011) 34–41.
308. Green, I.X., Tang, W., Neurock, M., Yates, J.T., Angew. Chem. Int. Ed.,
50 (2011) 10186–10189.
309. Golunski, S., Rajaram, R., Hodge, N., Hutchings, G.J., Kiely, C.J.,
Catal. Today, 72 (2002) 107–113.
310. Carrettin, S., Concepción, P., Corma, A., López-Nieto, J.M., Puntes,
V.F., Angew. Chem. Int. Ed., 43 (2004) 2538–2540.
311. Venezia, A.M., Pantaleo, G., Longo, A., Di Carlo, G., Casaletto, M.P.,
Liotta, F.L., Deganello, G., J. Phys. Chem. B, 109 (2005) 2821–2827.

b1469_Ch-02.indd 133 4/8/2013 12:26:21 PM


b1469 Catalysis by Ceria and Related Materials

134 J. J. Delgado et al.

312. Arena, F., Famulari, P., Trunfio, G., Bonura, G., Frusteri, F., Spadaro, L.,
Appl. Catal. B, Environ., 66 (2006) 81–91.
313. Lai, S.Y., Qiu, Y., Wang, S., J. Catal., 237 (2006) 303–313.
314. Pillai, U.R., Deevi, S., Appl. Catal. A, Gen., 299 (2006) 266–273.
315. Lu, J.-L., Gao, H.-J., Shaikhutdinov, S., Freund, H.J., Catal. Lett., 114
(2007) 8–16.
316. Shapovalov, V., Metiu, H., J. Catal., 245 (2007) 205–214.
317. Tang, Z.R., Edwards, J.K., Bartley, J.K., Taylor, S.H., Carley, A.F.,
Herzing, A.A., Kiely, C.J., Hutchings, G.J., J. Catal., 249 (2007)
208–219.
318. Wang, S.P., Zhang, T.Y., Wang, X.Y., Zhang, S.M., Wang, S.R., Huang,
W.P., Wu, S.H., J. Molec. Catal. A, Chem., 272 (2007) 45–52.
319. Widmann, D., Leppelt, R., Behm, R.J., J. Catal., 251 (2007) 437–442.
320. Dobrosz-Gómez, I., Kocemba, I., Rynkowski, J.M., Appl. Catal. B,
Environ., 83 (2008) 240–255.
321. Aguilar-Guerrero, V., Lobo-Lapidus, R.J., Gates, B.C., J. Phys. Chem. C,
113 (2009) 3259–3269.
322. Dobrosz-Gómez, I., Kocemba, I., Rynkowski, J.M., Catal. Lett., 128
(2009) 297–306.
323. Dobrosz-Gómez, I., Kocemba, I., Rynkowski, J.M., Appl. Catal. B,
Environ., 88 (2009) 83–97.
324. Huang, X.S., Sun, H., Wang, L.C., Liu, Y.M., Fan, K.N., Cao, Y., Appl.
Catal. B, Environ., 90 (2009) 224–232.
325. Li, Q., Zhang, Y., Chen, G., Fan, J., Lan, H., Yang, Y., J. Catal., 273
(2010) 167–176.
326. Liotta, L.F., Di Carlo, G., Pantaleo, G., Venezia, A.M., Catal. Today, 158
(2010) 56–62.
327. Shimada, S., Takei, T., Akita, T., Takeda, S., Haruta, M., Stud. Surf. Sci.
Catal., 175 (2010) 843–847.
328. del Río, E., Blanco, G., Collins, S.E., López-Haro, M., Chen, X.,
Delgado, J.J., Calvino, J.J., Bernal, S., Top. Catal., 54 (2011)
931–940.
329. Fu, Q., Weber, A., Flytzani-Stephanopoulos, M., Catal. Lett., 77 (2001)
87–95.
330. Fu, Q., Kudriavtseva, S., Saltsburg, H., Flytzani-Stephanopoulos, M.,
Chem. Eng. J., 93 (2003) 41–53.

b1469_Ch-02.indd 134 4/8/2013 12:26:21 PM


b1469 Catalysis by Ceria and Related Materials

Understanding Ceria-Based Catalytic Materials 135

331. Jacobs, G., Chenu, E., Patterson, P.M., Williams, L., Sparks, D.,
Thomas, G., Davis, B.H., Appl. Catal. A, Gen., 258 (2004) 203–214.
332. Liu, Z.-P., Jenkins, S., King, D.A., Phys. Rev. Lett., 94 (2005) 196102-1–4.
333. Deng, W., Jesus, J.D., Saltsburg, H., Flytzani-Stephanopoulos, M., Appl.
Catal. A, Gen., 291 (2005) 126–135.
334. Fu, Q., Deng, W., Saltsburg, H., Flytzani-Stephanopoulos, M., Appl.
Catal. B, Environ., 56 (2005) 57–68.
335. Deng, W., Flytzani-Stephanopoulos, M., Angew. Chem. Int. Ed., 45
(2006) 2285–2289.
336. Jacobs, G., Ricote, S., Davis, B.H., Appl. Catal. A, Gen., 302 (2006)
14–21.
337. Kim, C.H., Thompson, L.T., J. Catal., 244 (2006) 248–250.
338. Leppelt, R., Schumacher, B., Plzak, V., Kinne, M., Behm, R.J., J. Catal.,
244 (2006) 137–152.
339. Amieiro-Fonseca, A., Fisher, J.M., Ozkaya, D., Shannon, M.D.,
Thompsett, D., Top. Catal., 44 (2007) 223–235.
340. Denkwitz, Y., Karpenko, A., Plzak, V., Leppelt, R., Schumacher, B.,
Behm, R.J., J. Catal., 246 (2007) 74–90.
341. Karpenko, A., Leppelt, R., Cai, J., Plzak, V., Chuvilin, A., Kaiser, U.,
Behm, R.J., J. Catal., 250 (2007) 139–150.
342. Karpenko, A., Leppelt, R., Plzak, V., Behm, R.J., J. Catal., 252 (2007)
231–242.
343. Karpenko, A., Denkwitz, Y., Plzak, V., Cai, J., Leppelt, R., Schumacher,
B., Behm, R.J., Catal. Lett., 116 (2007) 105–115.
344. Meunier, F.C., Reid, D., Goguet, A., Shekhtman, S., Hardacre, C.,
Burch, R., Deng, W., Flytzani-Stephanopoulos, M., J. Catal., 247 (2007)
277–287.
345. Meunier, F.C., Goguet, A., Hardacre, C., Burch, R., Thompsett, D.,
J. Catal., 252 (2007) 18–22.
346. Rodríguez, J.A., Liu, P., Hrbek, J., Evans, J., Pérez, M., Angew. Chem.
Int. Ed., 46 (2007) 1329–1332.
347. Daly, H., Ni, J., Thompsett, D., Meunier, F.C., J. Catal., 254 (2008)
238–243.
348. Andreeva, D., Ivanov, I., Ilieva, L., Abrashev, M.V., Zanella, R.,
Sobczak, J.W., Lisowski, W., Kantcheva, M., Avdeev, G., Petrov, K.,
Appl. Catal. A, Gen., 357 (2009) 159–169.

b1469_Ch-02.indd 135 4/8/2013 12:26:22 PM


b1469 Catalysis by Ceria and Related Materials

136 J. J. Delgado et al.

349. Abd El-Moemen, A., Karpenko, A., Denkwitz, Y., Behm, R.J., J. Power
Sources, 190 (2009) 64–75.
350. Boaro, M., Vicario, M., Llorca, J., de Leitenburg, C., Dolcetti, G.,
Trovarelli, A., Appl. Catal. B, Environ., 88 (2009) 272–282.
351. Bond, G., Gold Bull., 42 (2009) 337–342.
352. Andreeva, D., Kantcheva, M., Ivanov, I., Ilieva, L., Sobczak, J.W.,
Lisowski, W., Catal. Today, 158 (2010) 69–77.
353. Boisen, A., Janssens, T.V.W., Schumacher, N., Chorkendorff, I., Dahl,
S., J. Molec. Catal. A, Chem., 315 (2010) 163–170.
354. Daly, H., Goguet, A., Hardacre, C., Meunier, F.C., Pilasombat, R.,
Thompsett, D., J. Catal., 273 (2010) 257–265.
355. Wang, Y., Liang, S., Cao, A., Thompson, R.L., Veser, G., Appl. Catal. B,
Environ., 99 (2010) 89–95.
356. Rodriguez, J.A., Catal. Today, 160 (2011) 3–10.
357. Luengnaruemitchai, A., Osuwan, S., Gulari, E., Int. J. Hydrogen Energy,
29 (2004) 429–435.
358. Trimm, D.L., Appl. Catal. A, Gen., 296 (2005) 1–11.
359. Ilieva, L., Pantaleo, G., Ivanov, I., Maximova, A., Zanella, R., Kaszkur, Z.,
Venezia, A.M., Andreeva, D., Catal. Today, 158 (2010) 44–55.
360. Laguna, O.H., Romero Sarria, F., Centeno, M.A., Odriozola, J.A.,
J. Catal., 276 (2010) 360–370.
361. Yi, G., Yang, H., Li, B., Lin, H., Tanaka, K.I., Yuan, Y., Catal. Today, 157
(2010) 83–88.
362. Tauster, S.J., Fung, S.C., Garten, R.L., J. Am. Chem. Soc., 100 (1978)
170–175.
363. Bernal, S., Calvino, J.J., Cauqui, M.A., Gatica, J.M., Larese, C., Perez
Omil, J.A., Pintado, J.M., Catal. Today, 50 (1999) 175–206.
364. Gatica, J.M., Baker, R.T., Fornasiero, P., Bernal, S., Blanco, G., Kašpar,
J., J. Phys. Chem. B, 104 (2000) 4667–4672.
365. Gatica, J.M., Baker, R.T., Fornasiero, P., Bernal, S., Kašpar, J., J. Phys.
Chem. B, 105 (2001) 1191–1199.
366. Sault, A.G., Madix, R.J., Campbell, C.T., Surf. Sci., 169 (1986) 347–356.
367. Lisowski, E., Stobinski, L., Dus, R., Surf. Sci., 188 (1987) L735–L741.
368. Stobinski, L., Appl. Surf. Sci., 103 (1996) 503–508.
369. Sermon, P.A., Bond, G.C., Wells, P.B., J. Chem. Soc., Faraday T. 1, 75
(1979) 385–394.

b1469_Ch-02.indd 136 4/8/2013 12:26:22 PM


b1469 Catalysis by Ceria and Related Materials

Understanding Ceria-Based Catalytic Materials 137

370. Lin, S., Vannice, M.A., Catal. Lett., 10 (1991) 47–61.


371. Jia, J., Haraki, K., Kondo, J.N., Domen, K., Tamaru, K., J. Phys. Chem.
B, 104 (2000) 11153–11156.
372. Cavanagh, R.R., Yates, J.T.J., J. Catal., 68 (1981) 22–26.
373. Stobinski, L., Dus, R., Appl. Surf. Sci., 62 (1992) 77–82.
374. Bus, E., Prins, R., van Bokhoven, J.A., Catal. Commun., 8 (2007)
1397–1402.
375. Shimizu, K.I., Miyamoto, Y., Kawasaki, T., Tanji, T., Tai, Y., Satsuma, A.,
J. Phys. Chem. C, 113 (2009) 17803–17810.
376. Groszek, A.J., Lalik, E., Appl. Surf. Sci., 257 (2011) 3192–3195.
377. Binet, C., Badri, A., Lavalley, J.C., J. Phys. Chem., 98 (1994) 6392–6398.
378. Wang, X., Andrews, L., Angew. Chem. Int. Ed., 42 (2003) 5201–5206.
379. Carabineiro, S.A.C., Nieuwenhuys, B.E., Gold Bull., 42 (2009)
288–301.
380. Gottfried, J.M., Schmidt, K.J., Schroeder, S.L.M., Christmann, K., Surf.
Sci., 536 (2003) 206–224.
381. Piccolo, L., Loffreda, D., Aires, F.J.C.S., Deranlot, C., Jugnet, Y.,
Sautet, P., Bertolini, J.C., Surf. Sci., 566–568 (2004) 995–1000.
382. Kim, J., Samano, E., Koel, B.E., J. Phys. Chem. B, 110 (2006)
17512–17517.
383. Weststrate, C.J., Lundgren, E., Andersen, J.N., Rienks, E.D.L.,
Gluhoi, A.C., Bakker, J.W., Groot, I.M.N., Nieuwenhuys, B.E., Surf.
Sci., 603 (2009) 2152–2157.
384. Sterrer, M., Yulikov, M., Fischbach, E., Heyde, M., Rust, H.P.,
Pacchioni, G., Risse, T., Freund, H.J., Angew. Chem., Inter. Ed., 45
(2006) 2630–2632.
385. Sterrer, M., Yulikov, M., Risse, T., Freund, H.J., Carrasco, J., Illas, F., Di
Valentin, C., Giordano, L., Pacchioni, G., Angew. Chem., Inter. Ed., 45
(2006) 2633–2635.
386. Rim, K.T., Eom, D., Liu, L., Stolyarova, E., Raitano, J.M., Chan, S.W.,
Flytzani-Stephanopoulos, M., Flynn, G.W., J. Phys. Chem. C, 113 (2009)
10198–101205.
387. Yang, F., Kundu, S., Vidal, A.B., Graciani, J., Ramírez, P.J., Senanayake,
S.D., Stacchiola, D., Evans, J., Liu, P., Sanz, J.F., Rodriguez, J.A., Angew.
Chem. Int. Ed., 50 (2011) 10198–10202.

b1469_Ch-02.indd 137 4/8/2013 12:26:22 PM


b1469 Catalysis by Ceria and Related Materials

138 J. J. Delgado et al.

388. Mehmood, F., Kara, A., Rahman, T.S., Henry, C.R., Phys. Rev. B, 79
(2009) 075422.
389. Mavrikakis, M., Stoltze, P., Norskov, J.K., Catal. Lett., 64 (2000)
101–106.
390. Bianchi, C.L., Pirola, C., Ragaini, V., Catal. Commun., 7 (2006)
669–672.
391. Nolan, M., Grigoleit, S., Sayle, D.C., Parker, S.C., Watson, G.W., Surf.
Sci., 576 (2005) 217–229.
392. Loschen, C., Bromley, S.T., Neyman, K.M., Illas, F., J. Phys. Chem. C,
111 (2007) 10142–10145.
393. Tauster, S.J., Acc. Chem. Res., 20 (1987) 389–394.

b1469_Ch-02.indd 138 4/8/2013 12:26:22 PM


b1469 Catalysis by Ceria and Related Materials

CHAPTER 3

INVESTIGATION OF THE OXYGEN


STORAGE AND RELEASE KINETICS
OF MODEL AND COMMERCIAL
THREE-WAY CATALYTIC MATERIALS
BY TRANSIENT TECHNIQUES
Angelos M. Efstathiou and Stavroula Y. Christou
Heterogeneous Catalysis Laboratory, Chemistry Department,
University of Cyprus, University Campus, P.O. Box 20537,
1678 Nicosia, Cyprus

3.1 Scope
The use of powerful analytical methodologies for the evaluation of
the oxygen storage capacity (OSC) and essential parameters of the
transient kinetics of oxygen storage and release in model and com-
mercial three-way catalysts (TWCs) are described in this chapter.
Sequences of alternate pulses or step changes in the feed-gas com-
position made at the inlet of a catalytic microreactor, temperature-
programmed, and transient isothermal 16O/18O isotopic exchange
methodologies are reviewed, with an emphasis on their power to
efficiently characterize oxygen mobility in CexZr1−xO2-based oxygen
storage components presently used in modern TWCs. An attempt is
made to describe the essential theoretical and experimental features
of these techniques through several examples. Case studies of the
deterioration of the oxygen storage and release kinetics in CeO2 and
CexZr1−xO2 solids due to chemical poisoning and thermal aging are

139

b1469_Ch-03.indd 139 4/8/2013 12:29:01 PM


b1469 Catalysis by Ceria and Related Materials

140 A. M. Efstathiou and S. Y. Christou

provided based on a variety of past and recent experimental investi-


gations. In this chapter, the regeneration of the activity of aged com-
mercial and model three-way catalysts through thermo-chemical
and acid-washing treatments, and how these are related to the recov-
ery of OSC and oxygen mobility, are also reviewed. The latter pro-
vides fundamental and important details on the mechanisms of
TWC deactivation in an effort to extend the life cycle of commercial
three-way catalysts by providing an improved design for their chemi-
cal composition and structural properties.

3.2 Introduction
It is well known that automotive catalytic converters have a highly
dynamic operation due to fluctuations in composition, tempera-
ture, and flow rate of an engine’s exhaust gases.1–3 Because of
this, much research has been conducted toward identifying the
chemical processes that determine the efficiency of TWCs under
cycling conditions of the air/fuel ratio in the frequency range of
0.1 Hz (rapid changes in traffic and driving conditions) to about
1 Hz (electronic control of the air/fuel ratio).1–5 In order to
understand and best describe the behavior of TWCs under the
oscillatory conditions in which they operate, it is imperative that
their characterization is carried out under such dynamic
conditions.
In order to reduce the effects of the prevailing oscillatory behav-
ior in an exhaust gas on the performance of a TWC, one of the
required properties for the support structure of the noble metals
used is its ability to store oxygen under “lean” conditions (excess
oxygen in the feed) and release it under “rich” conditions (oxygen
deficiency in the feed). This allows the TWC to operate in a wide
range of air/fuel ratios (λ) under dynamic conditions.6–12 This prop-
erty is termed the oxygen storage capacity (OSC), and it is a key
parameter for the appropriate performance of a TWC, namely, oxi-
dation of CO and hydrocarbons when the engine’s exhaust-gas com-
position is rich in oxygen, and reduction of NOx when the engine’s
exhaust-gas composition is on the lean side of the stoichiometric

b1469_Ch-03.indd 140 4/8/2013 12:29:02 PM


b1469 Catalysis by Ceria and Related Materials

Investigation of the Oxygen Storage and Release Kinetics of TWCs 141

value (λ=1).6–12 Various rare-earth metal oxides were initially widely


investigated as structural and electronic promoters of TWCs for
enhancing their activity, selectivity, and durability.13 The most impor-
tant material among these metal oxides used in three-way catalytic
chemistry was undoubtedly proven to be cerium oxide (CeO2),
because of its facile switching between Ce4+ and Ce3+ oxidation
states.6–13
As a result of the stringent automotive tailpipe emission regula-
tions, fuel combustion has to be precisely controlled, and as a
result of this, higher gas emission temperatures are obtained.
High-temperature sintering and deterioration of the OSC of CeO2
after extended use was found to lead to a deterioration of the
TWC’s performance, necessitating the production of improved
TWCs with increased performance durability and thermal stability.
This problem was solved through the insertion of smaller Zr4+ ions
into the ceria lattice, leading to a new generation of ceria–zirconia
solid solution (CexZr1−xO2−δ) OSC materials.13–22 CexZr1−xO2-based
materials are the state-of-the-art oxygen storage and release (OSR)
class of materials used in TWCs. Though many attempts to under-
stand the basic properties of these materials have been made, the
reasons for their good heat resistance and high OSC are still not
very well understood. One of the main reasons for this is that stud-
ies were performed in a fragmentary and not always quantitative
way. In particular, studies estimating the transient rate of oxygen
uptake and release under real driving conditions were rarely per-
formed, in spite of the importance of gas emission control. In a
modern vehicle, the air/fuel ratio λ is controlled by a cascade
feedback control algorithm with one λ sensor upstream, and a
second sensor downstream of the catalyst.23 The operation of the
controller, and especially the behavior of the second λ sensor are
strongly influenced by the OSC dynamics.23 Based on this rationale,
it is imperative to understand the oxygen storage and release prop-
erties of three-way catalytic materials. In particular, comprehensive
studies that quantitatively describe the transient kinetics of oxygen
storage and release via the participation of bulk and surface oxy-
gen diffusion processes as a function of fundamental properties of

b1469_Ch-03.indd 141 4/8/2013 12:29:02 PM


b1469 Catalysis by Ceria and Related Materials

142 A. M. Efstathiou and S. Y. Christou

the OSC material are highly desirable. This fundamental knowl-


edge would permit the design of the next generation of advanced
OSC materials.
Pulse and step-change transient OSC measurements, as well as
transient 18O-isotopic techniques, can provide the means to effec-
tively characterize commercial and model TWC materials for their
intrinsic oxygen storage and release kinetics under dynamic condi-
tions. Values for the most reactive (labile) oxygen and the total reac-
tive oxygen referred to as OSC and OSCC, respectively, of a given
TWC material are usually estimated after applying alternating pulses
and step changes to the feed-gas composition. Transient isothermal
isotopic exchange (TIIE) and temperature-programmed isotopic
exchange (TPIE) experiments using 18O2 gas are used to estimate
the transient rates and apparent activation energy barriers of both
surface and bulk oxygen diffusion during the transient oxygen iso-
topic exchange of gaseous 18O2 with surface/bulk lattice 16O.24–26
These techniques, when coupled with appropriate kinetic and math-
ematical modeling of the transient responses observed, are used to
determine the surface and bulk oxygen intrinsic diffusion coeffi-
cients, the concentration (µmol O·g−1), and the equivalent number
of monolayers of exchangeable oxygen species.24–26

3.3 Theoretical Background of Transient Techniques


Used in Oxygen Storage and Release Kinetic Studies
The ceria-based oxygen storage components of a TWC are capa-
ble of undergoing a fast change in the oxidation state of cerium
(Ce4+ ↔ Ce3+) after a change in exhaust-gas composition between
fuel-lean and fuel-rich conditions. Because of this change, the
ceria is continuously subjected to rapid reduction and oxidation
cycles, and its composition alternates between CeO2 and CeO2−x
according to the following reaction:13–22

Æ CeO2-x + x O2
H2 /CO
æææææ
CeO2 ¨ææææ æ (3.1)
H2O /CO2 2

b1469_Ch-03.indd 142 4/8/2013 12:29:02 PM


b1469 Catalysis by Ceria and Related Materials

Investigation of the Oxygen Storage and Release Kinetics of TWCs 143

Equation (3.1) can also be written in terms of ceria defect chemistry


as follows:

(3.2)

Equation (3.2) also applies to the creation of oxygen vacancies


at high temperatures in a vacuum or in the presence of an inert gas.
For the use of CO as a reductant or H2O as an oxidizing species,
Eq. (3.1), the following elementary reaction steps, Eqs. (3.3) and
(3.4), respectively, are appropriate:

2 CexCe + OxO + CO(g) ↔ 2 Ce′Ce + Vö + CO2(g) (3.3)

2 Ce′Ce + H2O(g) ↔ 2 CexCe + OxO + H2(g) (3.4)

3.3.1 Pulse-injection technique


Yao and Yu Yao6 first proposed the use of a pulse-injection method
for the measurement of the oxygen storage capacity of the CeO2-
containing catalytic materials used in TWCs under dynamic condi-
tions with an alternating pulse perturbation between oxidizing and
reducing gas atmospheres. This methodology of an alternate pulsing
between oxidizing and reducing gas atmospheres with a time domain
from about 1 s per pulse (∼0.5 Hz) up to several seconds was later
adapted by many researchers who were studying the oxygen storage
and release kinetics of ceria-based materials under experimental
conditions that simulate vehicle air/fuel modulation around the stoi-
chiometric λ value.7–12 According to this method of measurement,
Yao and Yu Yao6 proposed two different oxygen storage capacities:

(1) The OSC of a solid is defined as the concentration (µmol O·g−1)


of the most reactive and readily available (labile) oxygen atoms

b1469_Ch-03.indd 143 4/8/2013 12:29:02 PM


b1469 Catalysis by Ceria and Related Materials

144 A. M. Efstathiou and S. Y. Christou

Table 3.1 Sequence of feed-gas composition switches during a pulse-injection


experiment to measure OSC and OSCC.
Measurement Sequence of step changes in gas composition over
the catalyst sample
OSC 20 vol% O2/He (TOSC, 1 h) → He (TOSC, ∆t a) → one H2 or
CO pulse (e.g., 50 µmol/pulse) at TOSC → He (TOSC, ∆t a) →
successive oxygen pulses (e.g., 20 µmol O/pulse) until
saturation at TOSC
OSCC 20 vol% O2/He (TOSC, 1 h) → He (TOSC, ∆t a) → successive
pulses of H2 or CO at TOSC (e.g., 50 µmol/pulse) until no
uptake of H2 or CO is observed → He (TOSC, ∆t a) → successive
oxygen pulses (e.g., 20 µmol O/pulse) until saturation at TOSC
Note: a ∆t is the time needed to purge the gas lines and reactor of the O2, CO, or
H2 gas. Its value depends on the gas-flow system (Fig. 3.1), reactor volume, and
the gas flow rate (e.g., ∆t = 2–10 min).

in the solid. The OSC measured over a pre-oxidized sample


using CO or H2 as reductive species is defined as the amount of
CO2 or H2O (µmol·g−1) produced during the first pulse of CO
or H2, respectively. It can also be estimated from the CO or H2
uptake (µmol·g−1), or the amount of oxygen consumed (µmol
O·g−1) during the re-oxidation stage until saturation, following
the CO or H2 pulse (Table 3.1).
(2) The oxygen storage capacity complete (OSCC) of a solid is
defined as the maximum concentration (µmol O·g−1) of oxygen
species removed from the solid under the prevailing experimen-
tal conditions for the measurement (e.g., pulse size, T). The
OSCC measured over a pre-oxidized sample is the amount of
CO2 (µmol CO2·g−1) or H2O (µmol H2O·g−1) produced upon
introducing several consecutive pulses of CO or H2, respectively,
until no further CO or H2 consumption is recorded. It can also
be estimated from the amount of CO or H2 uptake (µmol·g−1),
or the concentration of oxygen consumed (µmol O·g−1) during
the re-oxidation stage until saturation, following the CO or H2
pulses.

b1469_Ch-03.indd 144 4/8/2013 12:29:03 PM


b1469 Catalysis by Ceria and Related Materials

Investigation of the Oxygen Storage and Release Kinetics of TWCs 145

According to Eq. (3.1), oxygen storage is formally considered as


the amount of lattice oxygen released (left-to-right) or stored (right-
to-left) under the rich (net reducing)/lean (net oxidizing) excur-
sions of the air/fuel ratio.13–22,27–31 The value of x in Eq. (3.1)
depends on several variables, such as the temperature, λ-oscillations,
and exhaust-gas composition.23
The experimental apparatus first used in pulse-injection experi-
ments was described by Yao and Yu Yao,6 while schematic diagrams of
it were reported by Su et al.,7 Engler et al.10 and Duprez and Descorne.19
In a characteristic CO pulse-injection experiment, the catalyst sample
(typically 10–50 mg) is placed in a quartz tubular or U-tube-type
microreactor kept at a given reaction temperature. A carrier gas (e.g.,
He or Ar) is used at flow rates within the 30–200 cm3·min−1 range to
sweep the pulse through the catalytic microreactor.
Figure 3.1 shows the chromatographic valve system needed for
conducting CO/O2 pulse-injection experiments.32 The V2 valve is
used to switch the feed gas from an oxidizing (e.g., 0.5–2.0 vol% O2/
He) to a reducing (e.g., 1.0–2.0 vol% CO/He) gas atmosphere. The
V5 valve is equipped with a gas sampling loop of known volume

Figure 3.1 An appropriate chromatographic valve system used for conducting


CO/O2 pulse-injection experiments for the measurement of OSC and OSCC. N.V.:
needle valve; B.M.: buble-meter; V2: 4-way chromatographic valve; V3, V5: 6-way
chromatographic valve.

b1469_Ch-03.indd 145 4/8/2013 12:29:03 PM


b1469 Catalysis by Ceria and Related Materials

146 A. M. Efstathiou and S. Y. Christou

(e.g., 0.5–1.0 mL) and a valve actuator operated so that the pulse
sequence and timing can be changed in an accurate and reproduc-
ible manner producing distinctive concentration pulse curves. The
gas loop is used to store the quantity (mol) of the oxidizing or
reducing gas to be injected into the He carrier gas stream, and then
to the reactor through the use of a six-way chromatographic valve
(V3). The exit stream from the microreactor is connected to an
online mass spectrometer for gas analysis. Table 3.1 shows the neces-
sary sequence of feed-gas composition switches applied during a
pulse-injection experiment for the measurement of OSC and OSCC,
using CO or H2 as the reducing gas (e.g., Eq. (3.3)).
In the experimental set-up shown in Fig. 3.1, the OSC/OSCC
measurements using the most common reducing/oxidizing gas
pairs, e.g., CO/O2 or H2/O2, take place as follows:
The catalyst sample is initially pre-treated in an O2/He gas mixture
(e.g., 20 vol% O2) for 2 h at the maximum temperature for the
OSC/OSCC measurements in order to ensure that the catalyst is
fully oxidized at the prevailing treatment conditions. The experi-
mental protocol following this oxygen pre-treatment for the meas-
urement of OSC on a typical Pd/CexZr1−xO2−δ catalyst is schematically
illustrated in Fig. 3.2, and is described as follows:

Step 1: After oxygen pre-treatment of the catalyst, the latter is raised


to the temperature at which the OSC is to be measured
(TOSC, Table 3.1) under a He flow.
Step 2: The catalyst sample is further treated with several pulses of
O2 at TOSC until saturation. During this step, oxygen is stored
on both the Pd metal and the ceria–zirconia support.
Step 3: Following step 2, a He flow is used to flush the gas lines and
the reactor volume from gas-phase oxygen.
Step 4: During this step, one CO pulse (e.g., 50 µmol for OSC meas-
urement) or several consecutive CO pulses (for OSCC meas-
urement) are injected into the catalytic reactor until no
further CO consumption or CO2 production is observed.
During this step, several dynamic phenomena occur, namely:

b1469_Ch-03.indd 146 4/8/2013 12:29:03 PM


b1469 Catalysis by Ceria and Related Materials

Investigation of the Oxygen Storage and Release Kinetics of TWCs 147

Figure 3.2 Schematic representation of the chemical effects on a Pd/CexZr1−xO2−δ


catalyst during the sequence of steps 1–4 applied during a CO pulse-injection
experiment for OSC measurement.

bulk/surface oxygen diffusion (oxygen back spillover), CO


chemisorption, and other surface reactions, which will be
discussed later.
Step 5: Following step 4, the gas lines and reactor volume are flushed
with a He flow to eliminate gas-phase oxygen.
Step 6: The catalyst sample is then treated with one O2 pulse or with
several consecutive O2 pulses, for OSC or OSCC, respec-
tively, until no further oxygen consumption is observed.
Re-oxidation of the catalyst sample takes place as in step 2.

OSC and OSCC are expressed in µmol O·g−1 as determined dur-


ing step 4 based on the CO consumption or CO2 production
dynamic responses (Fig. 3.3). However, this methodology may not

b1469_Ch-03.indd 147 4/8/2013 12:29:03 PM


b1469 Catalysis by Ceria and Related Materials

148 A. M. Efstathiou and S. Y. Christou

always give an accurate estimate of OSC/OSCC, as will be discussed


below. Alternatively, oxygen storage capacities can be measured dur-
ing step 6 using successive O2 pulses. During the latter pulse-oxygen
treatment, the consumption of oxygen is estimated. The latter
reflects both the amount of oxygen required to fill in the oxygen
vacancies, and that chemisorbed on the Pd metal or resulting in the
formation of subsurface PdO, which is usually significantly lower
than that for the CexZr1−xO2−δ support. A correction for PdO forma-
tion is possible via independent metal dispersion measurements. In
addition, any small amounts of CO or carbon accumulated during
the CO pulsing can be measured during the oxygen pulsing, and this
quantity is subtracted from the total amount of oxygen consumed.
The CO and CO2 gas-phase dynamic responses recorded during
step 4 of the CO pulse-injection experiment (OSC measurement)
are shown in Fig. 3.3. In the CO/He pulse, Ar gas of the same com-
position is added as a trace (e.g., 1 vol% CO/1 vol% Ar/He). The
Ar response represents the dynamics of a non-adsorbing and non-
reacting gas when pulsed (using the same amount as CO) through
the reactor with the catalyst present. The area difference between
the Ar and the CO response curves (shaded area in Fig. 3.3) is the
amount of CO consumed, or equivalently the OSC. In addition,
integration of the CO2 transient response gives the amount of CO2

Figure 3.3 Ar, CO, and CO2 gas-phase dynamic responses obtained during the
measurement of OSC in a CO pulse-injection experiment.

b1469_Ch-03.indd 148 4/8/2013 12:29:03 PM


b1469 Catalysis by Ceria and Related Materials

Investigation of the Oxygen Storage and Release Kinetics of TWCs 149

produced as the result of the interaction of CO with the pre-oxi-


dized catalyst surface.
The concentration (mol O·g−1) of oxygen species in the catalyst
required for the conversion of CO to CO2 (step 4, Fig. 3.2) is
obtained from the material balance, Eq. (3.5), based on the reac-
tion: CO (g) + OL → CO2 (g), where OL refers to lattice oxygen. If
no side reactions are considered, the difference in area between the
Ar and CO gaseous responses at the outlet of the reactor provides
the CO uptake (mol CO·g−1), or equivalently the OSC:32

t
FT f out
-1
OSC(mol O ◊ g ) = ◊ Ú ( y Ar - yCO
out
)dt (3.5)
W 0

where, FT is the molar flow rate of the He carrier gas (mol·s−1), W (g)
is the mass of the catalyst sample, and youtAr and youtCO are the mole
fractions of Ar and CO, respectively, at the outlet of the
microreactor.
The OSC and OSCC quantities strongly depend on the experi-
mental conditions applied, e.g., the nature and the partial pres-
sure of the reducing gas (H2 or CO), the pre-treatment and
reaction temperature (TOSC), the frequency of the pulses applied,
the chemical composition of the catalyst, and other structural
characteristics of the catalyst. Yao and Yu Yao6 first reported that
the OSC of CeO2, CeO2/Al2O3, and PM/CeO2–Al2O3 solids meas-
ured by the pulse-injection method was affected by the pre-treat-
ment temperature, the reaction temperature, the partial pressure
of CO, the presence and dispersion of precious metals (PM: Pt, Pd,
or Rh), and the concentration of CeO2 in the alumina matrix. For
example, the OSC was found to increase in the presence of noble
metals, with an increase in the partial pressure of the reducing gas,
and an increase in TOSC and the metal loading, whereas it was
found to decrease with an increase in the catalyst’s pre-treatment
temperature.
OSC and OSCC measurements are also performed using H2 as a
reducing agent,8,33–35 whereas hydrocarbons, like CH4, C3H6, and

b1469_Ch-03.indd 149 4/8/2013 12:29:03 PM


b1469 Catalysis by Ceria and Related Materials

150 A. M. Efstathiou and S. Y. Christou

C3H8, have rarely been used.8,36 It is very interesting though to exam-


ine the effect of these reducing agents on the OSC, since they are
all present in an exhaust-gas stream. Su and Rothschild8 found that
the efficiency of CO, H2, and C3H6 as reducing agents decreased in
the following order: H2 > CO > C3H6, whereas methane (CH4) was
found to provide about four times lower OSC values than CO.36
The reaction taking place during a CO pulse-injection experi-
ment for the measurement of OSC in ceria-based materials is given
by Eq. (3.3), whereas that during a H2 pulse-injection experiment is
given by:

H2 pulse: 2 CexCe + OxO + H2(g) ↔ 2 Ce′Ce + Vö + H2O(g) (3.6)

During the re-oxidation stage, the following reaction takes place:

O2 pulse : 0.5 O2(g) + 2 Ce′Ce + Vö → 2 CexCe + OxO (3.7)

A number of reactions might also take place during the measure-


ment of OSC in a TWC sample with CO/O2 pulses (Section 3.3.1)
or step gas concentration switches (Section 3.3.2). These are
described by the following equations:27,29,32,34,37–42

CO(g) + O − S1 → CO2 + S1 (3.8)

CO – S1 + O − S1 → CO2(g) + 2S1 (3.9)

O − S2 + S1 → O − S1 + S2 (3.10)

2 (S2) → CO 3 – S2
CO(g) + O2− 2−
(3.11)

CO2(g) + O2−(S2) → CO2−


3 − S2 (3.12)

2CO(g) + S1 → CO2(g) + C − S1 (3.13)

where S1 refers to a site present in the precious metal/metal oxide


(e.g., Pd/PdO), while S2 refers to a site present in the washcoat
(e.g., CeO2 or CexZr1−xO2−δ) of the TWC. During CO pulsing, at

b1469_Ch-03.indd 150 4/8/2013 12:29:03 PM


b1469 Catalysis by Ceria and Related Materials

Investigation of the Oxygen Storage and Release Kinetics of TWCs 151

relatively high temperatures, gaseous CO reacts with O atoms in the


pre-oxidized supported noble-metal nanocrystallites to form CO2 via
the Eley–Rideal mechanism, Eq. (3.8).32,37 Reaction (3.8) leads to the
partial reduction of the noble metal-oxide phase by forming an oxygen
vacancy. On partially reduced noble-metal nanoparticles, the
Langmuir–Hinshelwood mechanism may also take place through the
interaction between adsorbed O–S1 and CO–S1 species, Eq. (3.9).38
The formation of oxygen vacancies enables O atoms originally stored
in the support solid phases (e.g., CeO2 and CexZr1−xO2−δ) to diffuse
towards the metal/metal oxide inter-phase via a back-spillover mecha-
nism, Eq. (3.10).32,37 If only reactions (3.8)–(3.10) take place, the
amount of CO consumed is expected to be the same as the amount
of CO2 produced based on a carbon material balance. If this is not the
case, then several other reaction paths must be considered. For exam-
ple, CO or CO2 may react with surface lattice oxygen or adsorbed
dioxygen species (O22−) on the support phase(s) of the washcoat to
form various kinds of carbonate species, Eqs. (3.11) and (3.12). The
latter species are strongly bound on the support, and desorb at ele-
vated temperatures.39,40 Additionally, CO may disproportionate on the
noble metal to form CO2 (g) and adsorbed C–S1 species via the
Boudouard reaction, Eq. (3.13).39–42 Holmgren et al.41 reported that
with an increasing degree of reduction of the Pt/CeO2 catalyst, CO
disproportionation occurs to an increasing extent. It is appropriate to
point out that the extent of each of the reactions (3.8)–(3.13) occur-
ring during the CO pulse-injection experiment is determined by the
transient kinetics associated with each reaction. This transient kinetics
is influenced by the chemical state of the catalyst surface, the concen-
tration of CO in the gas phase, the reaction temperature, and the
initial amount of stored oxygen in the solid catalyst sample.
Reactions (3.11)–(3.13) to a significant extent limit the use of
CO as an appropriate reducing agent in a pulse-injection experi-
ment for determining the OSC in ceria-containing solids, particu-
larly in the presence of noble metals, because they cause large
differences in CO uptake and CO2 production, depending on the
reaction temperature.27,40–42 However, if the CO2 does not chemisorb
on an oxygen vacant site, and C–S1 does not react with oxygen, or

b1469_Ch-03.indd 151 4/8/2013 12:29:03 PM


b1469 Catalysis by Ceria and Related Materials

152 A. M. Efstathiou and S. Y. Christou

the latter chemical steps are limited to a small degree, then estima-
tion of the OSC based on the re-oxidation step 6 can be considered
as a reliable measurement.
Bernal et al.43 pointed out also the particular issues that need to
be considered in using H2 pulse-injection experiments for accurate
OSC measurements, for example, the hydrogen chemisorbed spe-
cies formed on the noble metal and those remaining after the He
purge (step 3, Fig. 3.2) prior to re-oxidation. The relative contribu-
tion of this chemisorbed hydrogen to the total oxygen consumption
may be important for high surface area supports and moderate
reduction temperatures (TOSC < 500 °C).43

3.3.2 Dynamic oxygen storage capacity (DOSC)


An estimation of the concentration (µmol O·g−1) of oxygen stored
or released under dynamic conditions that more closely simulate
the lean/rich fluctuating conditions in exhaust-gas composition
during real TWC operation (λ and frequency) is certainly much
more useful and relevant in correlating OSC with the activity of a
TWC.23 The concentration of oxygen stored or released, which is
kinetically available during the fast oscillations in the redox nature
of the exhaust gas in motor vehicles, is termed the dynamic oxygen
storage capacity (DOSC, µmol O·g−1).
The measurement of DOSC is typically carried out isothermally
by switching between reducing (CO/Ar/He) and oxidizing (O2/
Ar/He) gas atmospheres over a TWC (powder or monolith) sample.
The frequency (the number of step changes per second) of switch-
ing is within the 0.05–1.0 Hz range.29,34,37,44–51 The experimental pro-
cedure usually applied is that of alternating step changes in the
concentration of the feed-gas as shown in Table 3.2. At first, the cata-
lyst sample is initially pre-treated with 20 vol% O2/He gas mixture
for 1 h at the temperature of DOSC measurement (TDOSC). Following
a He purge at the same temperature, the feed is changed to an
x vol% O2/He gas mixture (x∼1–2) for a given period of time,
t (sec). A switch to He (TDOSC, ∆t) is then made to purge the oxygen
reactant gas and avoid a direct reaction between CO (the next step

b1469_Ch-03.indd 152 4/8/2013 12:29:03 PM


b1469 Catalysis by Ceria and Related Materials

Investigation of the Oxygen Storage and Release Kinetics of TWCs 153

Table 3.2 Step changes in gas concentration for DOSC measurements over a TWC
sample (FT = 30 cm3·min−1).34
Experiment Sequence of step changes in gas concentration
A 1.5% O2/He (TDOSC, t) → He (TDOSC, ∆t) → 3% CO/3% Ar/He
(TDOSC, t) → He (TDOSC, ∆t) → repeat cycle
B 10% CO2/1.5% O2/He (TDOSC, t) → He (TDOSC, ∆t) → 3% CO/3%
Ar/He (TDOSC, t) → He (TDOSC, ∆t) → repeat cycle
C 20 ppm SO2/1.5% O2/He (TDOSC, t) → He (TDOSC, ∆t) →
3% CO/3% Ar/He (TDOSC, t) → He (TDOSC, ∆t) → repeat cycle

in feed gas composition) and any gas-phase O2 left in the reactor. A


switch to a y vol% CO/Ar/He gas mixture (y∼1–4) at TDOSC is then
made, and the same procedure is repeated several times until repro-
ducible transient behavior is observed (Experiment A, Table 3.2).
The evolution of the gas-phase composition in the reactor is con-
tinuously monitored by an online mass spectrometer, which records
the Ar (m/z = 40), CO (m/z = 28), CO2 (m/z = 44), and O2 (m/z = 32)
signals. The second step in the experimental procedure in Table 3.2
is used for the quantitative measurement of DOSC.
Figure 3.4 shows the transient response curves of CO, CO2, and
Ar obtained during the 3 vol% CO/3 vol% Ar/94 vol% He step in
the gas delivery sequence of Experiment A (Table 3.2) at 400 °C
performed over a commercial Pd–Rh TWC.34 The presence of Ar in
the reducing gas mixture enables the simultaneous recording of the
transient response of a non-adsorbing and non-reacting gas, which
passes from the switching valve through the reactor (with the cata-
lyst bed in place) to the mass spectrometer. The recording of this Ar
response signal is necessary in order to evaluate correctly the con-
sumption of CO (µmol CO.g−1), which is based on the area between
the Ar and CO response curves. The DOSC can be also quantified
based on CO2 production by integrating the CO2 response obtained
during a single O2/CO cycle (Fig. 3.4).
In general, the CO2 profile has two peaks during a DOSC exper-
iment. The first peak arises during the CO step and the second one
during the O2 step. This is illustrated in Fig. 3.5, where the evolution

b1469_Ch-03.indd 153 4/8/2013 12:29:03 PM


b1469 Catalysis by Ceria and Related Materials

154 A. M. Efstathiou and S. Y. Christou

3%CO/3%Ar/He
3
Ar

Concentration (mol %) 2 CO
TDOSC=400oC

1 CO2

0
0 5 10 15 20 25 30
Time (sec)

Figure 3.4 Transient response curves of CO, CO2, and Ar during the CO step of
the gas delivery sequence: 1.5% O2/He (30 s) → He (30 s) → 3% CO/3% Ar/He
(t) at 400 °C over a commercial Pd–Rh TWC. Adapted with permission from
Lambrou et al.34 Copyright 2004 Elsevier.

Figure 3.5 Partial pressures of CO (- - -), O2 (…), and CO2 (—) during the gas
delivery sequence: 4% CO/He (5 s) → He (5 s) → 2% O2/He (5 s) over CeO2 at
400 °C. Adapted with permission from Boaro et al.29 Copyright 2004 Elsevier.

of CO, O2, and CO2 partial pressures as a function of time during


the gas delivery sequence 4% CO/He (5 s) → He (5 s) → 2% O2/He
(5 s) applied over CeO2 at 400 °C is shown.29 Similar behavior was
observed in other studies.45,48,52 During the first half of the period

b1469_Ch-03.indd 154 4/8/2013 12:29:03 PM


b1469 Catalysis by Ceria and Related Materials

Investigation of the Oxygen Storage and Release Kinetics of TWCs 155

(CO step) CO2 production (peak 1) takes place through the reac-
tion of CO with oxygen stored in the solid sample (CeO2−δ) and the
CO2 (1) peak is in phase with the CO pulse. The O2 peak is delayed
compared to the CO2 (2) peak (Fig. 3.5) because oxygen is first
utilized by ceria for re-oxidation.29 The interpretation of the appear-
ance of CO2 (peak 2) in the second half of the period (O2 step) is
not straightforward. Boaro et al.29 were among the first to investigate
in some detail the origin of this CO2 (2) peak. After performing in
situ diffuse reflectance Fourier transform spectroscopy–mass spec-
trometry (DRIFTS-MS) studies during the same CO/O2 DOSC
experiment, they concluded the following: (i) when CO is pulsed
into the reactor, the ceria support is reduced and CO2 forms; (ii) a
fraction of this CO2 is instantaneously released, thus giving rise to
the CO2 (1) peak (Fig. 3.5). The remaining CO2 strongly interacts
with the ceria support, forming a carbonate-like species; (iii) during
the oxygen pulse, the carbonate-like intermediates decompose to
release CO2 (peak 2), or alternatively, oxygen is adsorbed on oxygen
vacant sites (reduced Ce3+ sites), thus re-oxidizing the ceria and
releasing CO2. It is also likely that in other metal-oxide OSC materi-
als or supported noble-metal catalysts, the CO2 (2) peak represents
the reaction of oxygen with carbon, the latter originating from CO
disproportionation, Eq. (3.13), during the first half cycle (CO puls-
ing). For supported noble metals, it is also possible that during the
He purge, following the CO step, some strongly adsorbed CO on
the noble metal remains, which then readily reacts with oxygen dur-
ing the O2 step to form CO2.
As mentioned earlier, the DOSC strongly depends on the lean/
rich engine exhaust-gas composition.23 Lambrou et al.34 investigated
the effects of the presence of 10 vol% CO2 or 20 ppm SO2 in the
oxidizing gas mixture on the DOSC (experiments B and C, respec-
tively, in Table 3.2). The concentrations of CO2 and SO2 used were
similar to those encountered in the exhaust-gas composition of a
gasoline-driven car. It was found that the amount of CO2 produced
(Fig. 3.4) decreases in the presence of 10% CO2 or 20 ppm SO2 in
the O2/He feed-gas stream. It was suggested that the presence of
CO2 or SO2 in the oxidizing gas mixture, under which oxygen is

b1469_Ch-03.indd 155 4/8/2013 12:29:04 PM


b1469 Catalysis by Ceria and Related Materials

156 A. M. Efstathiou and S. Y. Christou

stored, does not affect the quality of sites associated with the back
spillover of labile oxygen species, but only the concentration of
stored oxygen due to the formation of stable carbonate or sulfate
species in ceria.
Boaro et al.44 and Zhao et al.48 investigated the rate of CO oxida-
tion under steady-state reaction conditions (4% CO/2% O2/He gas
mixture), and that obtained during OSC measurements by (i) pulse
injection (one 4% CO/He pulse followed by five 2% O2/He pulses),
and (ii) alternate switching 4% CO/He (1 s) → He (2 s) → 2% O2/
He (1 s). It was found that the rate of CO oxidation under stationary
conditions over a series of CexZr1−xO2−δ (0 ≤ x ≤1) solids is influ-
enced by the number of exposed surface redox sites of Ce4+/Ce3+.
When the CO oxidation reaction was carried out under transient
conditions (OSC measurements), oxygen diffusion plays an impor-
tant role in the enhancement of the redox properties of ceria–
zirconia compared to ceria. It was found that the performance of
solid ceria–zirconia for CO oxidation was significantly better than
that of ceria, with optimum performance in the middle composi-
tional range, especially with low-surface area samples, for which
participation of lattice oxygen is more likely to occur.44,48 It is thus
clear that under dynamic conditions, the structural features of the
catalysts are more important in evaluating the OSC. The effect of
intrinsic microstructure variations in ceria–zirconia on Zr4+ coordi-
nation can be correlated to the oxygen mobility.53 This is the
primary reason why CexZr1−xO2−δ materials are used nowadays in
TWCs in place of pure CeO2.
OSC investigations on a time scale of a few seconds, such as those
previously discussed, are usually interconnected with gas-phase diffu-
sion (inter/intra particle) and desorption phenomena, the dynamics
of which is slower than surface phenomena. In addition, the actual
change in composition of the exhaust gas for a TWC between lean
and rich occurs on a millisecond scale. In this respect, only a limited
number of studies have presented OSC data on a millisecond
scale.45,54–56 The surface science approach56 employs an ultrahigh
vacuum with gas-pulsed valves, a catalyst on a planar substrate, and a
turn over frequency (TOF) mass spectrometer. This set-up allows

b1469_Ch-03.indd 156 4/8/2013 12:29:04 PM


b1469 Catalysis by Ceria and Related Materials

Investigation of the Oxygen Storage and Release Kinetics of TWCs 157

millisecond oxygen storage capacity (MS-OSC) measurements to be


successfully performed, after the elimination of diffusion problems.
These pulse-injection and step-change methodologies for esti-
mating OSC and DOSC are exclusively adopted in laboratory studies
using microreactors with small quantities of model or practical
three-way catalysts (in powder form or as pieces of monolith) and
simulating lean/rich exhaust-gas compositions. There is another
methodology for estimating the DOSC, which uses a bench engine
and other equipment.57 In this methodology, the oxygen storage
quantity (mol O2) in the whole TWC is represented as the integrator
of the deviation of the air/fuel ratio from the stoichiometric value
(14.6–14.7 w/w). Air/fuel data before and after the catalyst and
conversion efficiency are measured where the front-headed O2 sen-
sor (HO2S) feedback gain is high or low. When the gain is high, the
air/fuel perturbation is large, and when the gain is low, the pertur-
bation is small. This means that when the vehicle speed is constant,
the perturbation amplitude in high gain is so great that the catalyst
cannot absorb all the oxygen variation, and the rear air/fuel ratio
does not stay stoichiometric, thus deteriorating the efficiency of
conversion of the gas pollutants. When the vehicle speed is increas-
ing with low gain, the front air/fuel ratio in the transient state is out
of stoichiometric, which prevents the catalyst from absorbing all the
oxygen variation, and also the rear air/fuel ratio is out of stoichio-
metric. Avoiding the latter situation requires suitable air/fuel con-
trol in accordance with the OSC. In order to achieve this suitable
control, a relation between OSC, catalyst temperature, conversion
efficiency, and deterioration condition must be established.57

3.3.3 Modeling of OSC


For a complete understanding of the oxygen storage and release
features of model and practical TWCs containing ceria-based mate-
rials, a detailed kinetic model based on elementary reaction steps,
which describe the dynamics of all chemical and diffusion (surface/
bulk) processes, is imperative. Attempts at a kinetic model were
reported for the CO oxidation reaction after applying the pulse32

b1469_Ch-03.indd 157 4/8/2013 12:29:04 PM


b1469 Catalysis by Ceria and Related Materials

158 A. M. Efstathiou and S. Y. Christou

and step-change29,37,55,58,59 feed concentration methodologies over


model TWCs. A comprehensive OSC model was also incorporated
in the modeling of a real TWC dynamic operation.60 In a recent
work,23 the reversible nature of reduction of ceria by CO and H2 to
form CO2 and H2O, respectively, was proved to play an important
role in the oxygen storage/release dynamics based on kinetic mod-
eling of the observed transient gas-phase response concentrations.
In the same work,23 reference is made to previous work that consid-
ered oxygen storage related reactions in an effort to model three-
way catalytic chemistry.

3.3.4 Oxygen isotopic exchange


Surface and bulk diffusion processes of active species play a very
important role in catalysis not only in the catalytic process itself but
also during the activation and regeneration treatments of catalysts.
Oxygen storage and release in three-way catalysts and in many oxida-
tion reactions involving surface diffusion (spillover) of active oxy-
gen species are well-known examples where diffusion on the surface
and within the bulk of the solid largely influences the catalytic rate
and selectivity. However, detailed studies of their influence on the
intrinsic kinetic parameters are limited. 18O/16O isotopic exchange
has been proven to be a very powerful technique for studying the
kinetics of elementary steps involved in the oxygen storage and
release processes.

3.3.4.1 Mechanism of oxygen exchange


16
O/18O isotopic exchange has been studied over several metal
oxides24,61–73 and supported metal catalysts.25,26,74–83 According to
Boreskov,84 Winter,85 and Novakova86 there are three types of oxygen
exchange:

(1) Homoexchange or equilibration (Type I, R0, or R), which


occurs between dioxygen 18O2 and 16O2 molecules in the gas

b1469_Ch-03.indd 158 4/8/2013 12:29:04 PM


b1469 Catalysis by Ceria and Related Materials

Investigation of the Oxygen Storage and Release Kinetics of TWCs 159

phase without any participation of the 16O atoms present on the


surface of the solid oxide:

18
O2(g) +16O2(g)  218O16O(g)
(3.14)

(2) Simple heteroexchange (Type II, R1, or R′), which involves one
surface oxygen atom from the metal oxide, 16O(s), and one 18O
atom from the 18O2 isotope gas molecule according to the fol-
lowing steps:

18
O2(g) +16O(s)  18O16O(g)+18O(s) (3.15)

18
O16O(g) +16O(s)  16O2(g)+18O(s) (3.16)

(3) Multiple heteroexchange (Type III, R2, or R′′), which includes


two surface oxygen atoms from the oxide, 16O(s), and 18O2(g)
according to the following steps:

18
O2(g) + 216O(s)  16O2(g)+ 218O(s) (3.17)

18
O16 O(g) + 216 O(s)  16 O2(g)+16 O(s)+18 O(s) (3.18)

18
O16 O(g) + 216 O(s)  16 O2(g)+16 O(s)+18 O(s) (3.19)

Homoexchange experiments use an equimolar mixture of


16
O2(g) and 18O2(g) and monitor the formation of 18O16O(g),
Eq. (3.14), with time in order to study the oxygen activation on the
solid surface.19,25,26 On the other hand, in the heteroexchange
experiments, a mixture of 18O2(g) in He or Ar is passed over a solid
pre-oxidized with 16O2(g), and the production of 18O16O and 16O2
gases is monitored as a function of time (TIIE) or temperature
(TPIE). These experiments can be used to study surface and bulk

b1469_Ch-03.indd 159 4/8/2013 12:29:04 PM


b1469 Catalysis by Ceria and Related Materials

160 A. M. Efstathiou and S. Y. Christou

18
O16O 18
16 O2 18
O2 O2
6 1
6 16 18 8
16 O O
O 18
O
2 18 18 16
16 5 O O O
O 3 4
Metal particle

16
Metal oxide support O

Figure 3.6 Schematic representation of the elementary steps potentially involved


in 18O/16O isotopic exchange over a supported metal catalyst.

oxygen diffusion kinetics (Sections 3.3.4.2 and 3.3.4.3). As schemati-


cally shown in Fig. 3.6, after passing an 18O2/He gas mixture over a
supported metal catalyst that was initially pre-treated with an 16O2/
He gas mixture (heteroexchange experiment), the mechanism of
18
O/16O isotopic exchange includes the following steps:19,26,80

(1) adsorption and dissociation of 18O2 on the metal particles;


(2) spillover of 18O atoms from the metal to the support, as well
as back-spillover of 18O atoms from the support to the metal;
(3) diffusion of 18O atoms on the surface;
(4) exchange of 18O atoms with the 16O atoms of the support;
(5) back-spillover of 16O atoms from the support to the metal; and
(6) formation of 18O16O or 16O2 which desorb from the catalyst
surface.
At reaction temperatures higher than 400 °C, two more steps might
be involved:65
(7) bulk diffusion and exchange of 18O with the 16O atoms of the
support; and
(8) direct exchange of 18O2 from the gas phase with the 16O
atoms of the support.

Information about the exchange mechanism can be obtained


from the relative evolution of the oxygen isotopomer concentrations

b1469_Ch-03.indd 160 4/8/2013 12:29:04 PM


b1469 Catalysis by Ceria and Related Materials

Investigation of the Oxygen Storage and Release Kinetics of TWCs 161

at the beginning of the exchange process in the experiment. More


precisely, the production of 18O16O (m/z =34) primarily indicates that
the simple heteroexchange mechanism (Type II) is favored, whereas
the formation of 16O2 (m/z =32) reveals that the multiple heteroex-
change mechanism (Type III) prevails in the early stages of the
exchange. On simple metal oxides, such as ceria, alumina, zirconia,
magnesia, etc., oxygen exchange has been found to take place via a
simple exchange mechanism.24 In contrast, CexZr1−xO2−δ mixed metal
oxides predominately exchange their oxygen via a multiple exchange
mechanism.65
The type of exchange mechanism also provides information on
the nature of the mobile adsorbed oxygen species. In situ Fourier
transform infrared spectroscopy (FTIR) can be used to obtain infor-
mation about the nature and reactivity of surface dioxygen formed
on adsorption of 16O2 and the subsequent exchange of surface oxy-
gen species with gaseous 18O2. Adsorption and exchange experi-
ments are usually carried out at room temperature due to the very
short lifetime of the surface adsorbed oxygen species formed. The
exchange of these species with gas phase 18O2 is fast, and increases
with increasing 18O2 partial pressure. The formation of surface O−2
superoxide species is evident by the narrow band centered at
1126 cm−1 due to the O–O stretching vibrational mode, and its over-
tone at 2237 cm−1. The formation of superoxide species could be
seen as the transfer of one electron from a reduced cerium center
(Ce3+) of the metal oxide to the oxygen molecule, according to the
following reaction step:87

(Ce3+…)s + O2 (g) → (Ce4+…O−2)s (3.20)

Similarly, peroxide species O 2−


2 could be the result of the reaction
of molecular oxygen with two neighboring reduced cerium centers,
where two electrons are transferred from the solid to the molecular
oxygen:87

(Ce3+… …Ce3+)s + O2 (g) → (Ce4+…O2−


2

Ce4+)s (3.21)

b1469_Ch-03.indd 161 4/8/2013 12:29:04 PM


b1469 Catalysis by Ceria and Related Materials

162 A. M. Efstathiou and S. Y. Christou

Characteristic vibrations of 18O16O− and 18O−2 in CeO2 are


expected around 1094 and 1062 cm−1, respectively.88,89 The integral
band 1126 cm−1 normalized per unit of mass of solid can provide the
relative concentration of O −2 species. Duprez and co-workers66,90
observed that the population of surface superoxide species varied
with the composition of the CexZr1−xO2−δ solid solution. They found
that the concentration correlated with the oxygen storage capacity
measured by CO/O2 pulses. It was therefore suggested that such
binuclear oxygen species present on the surface of a mixed metal
oxide might be involved in oxygen transport and storage. In fact, the
most reactive surface dioxygen species (16O −2) adsorbed on ceria–
zirconia was shown to be exchanged selectively with 18O −2 via the
multiple exchange mechanism.66,90 These species can be seen either
as initiators or intermediates in the oxygen activation process on the
oxide surface (storage) prior to diffusion. Superoxides and perox-
ides have been postulated to be intermediate species in the com-
plete re-oxidation process of CeO2−x, and precursors of lattice
oxygen with electrons being progressively transferred from the solid
to the dioxygen molecule:66,87–89,91

-
O2(g) Æ O2ads Æ O2ads Æ O22-ads Æ 2O2-ads Æ 2O2- lattice (3.22)

3.3.4.2 Temperature-programmed isotopic exchange (TPIE)


A typical experimental procedure applied during TPIE measure-
ments is as follows (see Table 3.3). After calcination of the sample in
20 vol% 16O2/He at 850 °C for 2 h, followed by cooling in a He flow
to room temperature, the feed is switched to an x vol% 16O2/He gas
mixture for 15 min. A switch to the equivalent isotopic x vol% 18O2/
He gas mixture is then made, while at the same time the tempera-
ture of the catalyst is increased to 850 °C at the rate of 30 °C·min−1.
All three oxygen isotopic species, namely, 18O2 (m/z = 36), 18O16O
(m/z = 34), and 16O2 (m/z = 32), are continuously monitored by an
online mass spectrometer (MS) during the exchange process
(Fig. 3.7(a)). Typically, the sample mass used is 30–100 mg and the
flow rate of the oxygen gas mixture is 20–50 mL·min−1.

b1469_Ch-03.indd 162 4/8/2013 12:29:04 PM


b1469 Catalysis by Ceria and Related Materials

Investigation of the Oxygen Storage and Release Kinetics of TWCs 163

5%18O2/He 5%18O2/He
5.0 5.0
16
16
O2 O2
18
18 O16O

Concentration (mol%)
Concentration (mol%)

4.0 O2 4.0
18
18 16
O O O2

3.0
3.0
tmax T=700oC

2.0 2.0

1.0 1.0 ts.s.

0.0 0.0
200 300 400 500 600 700 800 T = const. 0 1000 2000 3000 4000 5000
Temperature ( oC) Time (s)
(a) (b)

Figure 3.7 Transient response curves of 16O2, 18O16O, and 16O2 gaseous species
obtained during (a) TPIE and (b) TIIE experiments over a CexZr1−xO2−δ solid.

3.3.4.3 Transient isothermal isotopic exchange (TIIE)


A typical experimental procedure applied during TIIE measure-
ments is as follows (see Table 3.3). After the sample’s pre-treatment
in 20 vol% 16O2/He gas mixture at 850 °C for 2 h, followed by cool-
ing in a He flow to the temperature of the exchange reaction (T),
the feed is switched to an x vol% 16O2/He gas mixture for 15 min.
A switch to the equivalent x vol% 18O2/He isotopic gas mixture at
the same flow rate is then made, while all three oxygen isotopic spe-
cies, 18O2, 18O16O, and 16O2, are continuously monitored with an
online mass spectrometer (Fig. 3.7(b)). TPIE and TIIE experiments
can be conducted using either a batch-recycled or a once-through
flow microreactor. The conversion of the various oxygen isotopic MS
signals into concentrations (mol%) is performed using 16O2/He and
18
O2/He calibration gas mixtures. Calibration of the 18O16O(g) MS
signal is usually based on the mean value of the known MS sensitivi-
ties obtained for the two other oxygen gas isotopes (16O2 and 18O2).

3.3.4.4 Description of surface and bulk oxygen diffusion during


TPIE and TIIE studies
The mathematical relations frequently used to determine rates (R),
diffusion coefficients (D), and the concentration of exchangeable

b1469_Ch-03.indd 163 4/8/2013 12:29:05 PM


b1469 Catalysis by Ceria and Related Materials

164 A. M. Efstathiou and S. Y. Christou

Table 3.3 Sequential changes in feed-gas composition in TPIE and TIIE


experiments.
Experiment Sequence of step changes in gas flow over the catalyst sample
TPIE 20 vol% O2/He (850 °C, 2 h) → cool down in He flow (25 °C) →
x vol% 16O2/He (25 °C, 15 min) → x vol% 18O2/He (25 °C) →
TPIE under x vol% 18O2/He flow up to 850 °C (e.g., use of
30 cm3·min−1 flow rate, and β = 30 °C.min−1 heating rate)
TIIE 20 vol% O2/He (850 °C, 2 h) → cool down in He flow to the
reaction temperature (T) → x vol% 16O2/He (T, 15 min) →
x vol% 18O2/He (T, ts.s.)

oxygen due to surface and bulk oxygen diffusion during TPIE and
TIIE experiments are described below.

Rate of exchange
The transient rate of 18O16O (g) production during an isothermal
step x vol% 16O2/He → x vol% 18O2/He in an isotopic exchange
experiment conducted in an open-flow microreactor can be esti-
mated based on a material balance given by the following
relationship:

FT
R16O18O (µmol16 O18 O ◊ g -1 ◊ s -1 ) = ◊ y16O18O (t ) (3.23)
W

where, y 18O16O (t) is the mole fraction of 18O16O(g) at a given time t


of the transient experiment, FT is the total molar flow rate (µmol·s−1)
of the isotopic gas mixture at the outlet of the reactor, and W is the
mass of catalyst sample used. It should be noted that in the material
balance described in Eq. (3.23), the accumulation term NG dy 18O16O/
dt is neglected; NG is the total number of moles in the gas-phase vol-
ume of the catalytic bed for a plug-flow reactor or in the gas-phase
volume of a continuous stirred tank reactor (CSTR) with the catalyst
in place. This material balance, Eq. (3.23) implies that external and
internal mass transport resistances are neglected within the catalytic

b1469_Ch-03.indd 164 4/8/2013 12:29:05 PM


b1469 Catalysis by Ceria and Related Materials

Investigation of the Oxygen Storage and Release Kinetics of TWCs 165

bed volume. The latter can be fulfilled according to well-established


criteria.92
Similarly, the transient rate of 16O2(g) formation during the
same isotopic exchange experiment can be estimated based on the
experimentally measured mole fraction y16O2 (t) given by the follow-
ing relationship:

FT
R16O (µmol16O2◊ g -1◊ s -1 ) = y16O (t ) (3.24)
2 W 2

16
The overall rate of exchange (R E) of O atoms from the solid is
given by the following relationship:

R E (µmol16O◊ g -1◊s -1 ) = 2R16O + R16O18O (3.25)


2

After integration of the R18O16O(t) curve with time using


Eq. (3.23), and the R16O2 (t) curve (Eq. (3.24)) with respect to the
Ar curve (gas switch 16O2/He/Ar → 18O2/He),93 the concentra-
tion of exchangeable 16O atoms, Q16O (µmol·g−1) in the solid at
any time during the exchange process can be estimated. Based
on this amount, the equivalent number of monolayers of
exchangeable oxygen, Ne, can be estimated by the following
relationship:

Q16O ◊ N A
Ne = (3.26)
SSA ◊ a

where, NA is Avogadro’s number (atom·mol−1), SSA is the specific


surface area of the solid (m2·g−1), and α is the number density of
surface oxygen atoms (atom·nm−2) of the given solid. For CeO2 and
CexZr1−xO2−δ, the average number density of surface oxygen atoms
assuming an equal distribution of the (100), (110), and (111) faces
was estimated to be between 13.1 and 14.4.65

b1469_Ch-03.indd 165 4/8/2013 12:29:05 PM


b1469 Catalysis by Ceria and Related Materials

166 A. M. Efstathiou and S. Y. Christou

Apparent activation energies for surface and bulk oxygen mobility


In a typical TPIE experiment, the equivalent number of monolayers
of 16O atoms exchanged can be calculated as a function of time, and
the time elapsed for which the first oxygen monolayer is exchanged
can be also determined. Prior to this time (Ne < 1), the apparent
activation energy for surface oxygen mobility (Esapp, kJ·mol−1) can
be obtained. This is achieved by using the overall rate of oxygen
exchange in the low-temperature range of the TPIE trace (Fig. 3.7(a))
to construct the Arrhenius plot (LnRE (t < tNe=1) versus 1/T). The
apparent activation energy for surface oxygen mobility obtained
from this procedure reflects surface adsorption, exchange, and dif-
fusion processes. Likewise, the apparent activation energy for sub-
surface and bulk oxygen diffusion (Ebapp, kJ·mol−1) can be estimated
using the high-temperature range of the TPIE profile (t > tN=1).73
It is expected that the measurement of R18O16O at t > tN=1 reflects to
a great extent the diffusion process of lattice 16O from the bulk of
the solid towards its surface, and to a lesser extent activated surface
mobility processes.
For the TIIE experiment (Fig. 3.7(b)), at the maximum concen-
tration shown in the 18O16O(g) response curve (tmax), the accumula-
tion term related to Eq. (3.23) is zero. The same is also true for the
pseudo-steady state of the transient (ts.s.). Therefore, no error is
max
introduced in estimating R18O16O through Eq. (3.23) and also the
apparent activation energy of surface oxygen mobility. Isothermal
exchange experiments can also be used to estimate the pseudo-
steady state rate and the apparent activation energy and diffusivity
s.s.
associated with the bulk oxygen diffusion process, using the R18 O16O
values obtained at different oxygen exchange temperatures.72
When surface exchange is the rate-determining step in the over-
all oxygen diffusion process, the surface diffusivity Ds (cm2·s−1) for
supported metal catalysts can be calculated based on a model previ-
ously developed by Kramer and Andre:94

Ê p ˆ Ê S1 ˆ
DS = Á ˜ ◊ Á
Ë 4 ¯ Ë C18 I o ¯˜
(3.27)
O

b1469_Ch-03.indd 166 4/8/2013 12:29:05 PM


b1469 Catalysis by Ceria and Related Materials

Investigation of the Oxygen Storage and Release Kinetics of TWCs 167

where, C18O is the concentration of 18O atoms (mol·cm−3), Io is the


specific perimeter (cm) of the metal particles considered as circular
sources of diffusing oxygen (Io = 2πrN), and S1 is the initial slope of
the curve representing the variation in the number of oxygen atoms
diffusing on the support surface at a given time of the exchange, N te,
with respect to the square root of time: S1 = dN et /d t . At any given
time of the exchange process, N te is given by:

2
N et = C18O I o Dst = 4rNC18O p Dst (3.28)
p

where, N and r are the number and radius of the metal particles,
respectively.
The effective diffusion coefficient for bulk oxygen diffusion
(D eff
b ) within the subsurface and bulk of the solid can be deter-
mined according to Peil et al.95 using the pseudo-steady state rate of
18 16
O O production at time equal to ts.s. (Fig. 3.7(b)):

2
ÊN ˆ
Deff
b = p t s.s. Á a ˜ (3.29)
Ë Cα O ¯

where Nα is the flux (mol·cm−2·s−1) of atomic oxygen from the bulk


to the surface, and Cαo (mol·cm−3) is the initial concentration of
atomic oxygen in the bulk of the solid. Nα is given by:

FT .
Na = y18O16O (s .s .) (3.30)
W ◊ SSA

where FT is the total molar flow rate (mol·s−1) of the 18O2/He gas
mixture used.
An advanced kinetic analysis of 18O/16O isotopic exchange over
CeO2–ZrO2–La2O3 and Pt/CeO2–ZrO2 solids was recently reported96
under conditions of dynamic adsorption-desorption equilibrium in
a plug-flow microreactor at 1 atm in the temperature range 650–850 °C

b1469_Ch-03.indd 167 4/8/2013 12:29:05 PM


b1469 Catalysis by Ceria and Related Materials

168 A. M. Efstathiou and S. Y. Christou

(SSITKA experiment), similar to the experiment of Fig. 3.7(b). In


this kinetic analysis, all three types of isotope exchange mechanism
(Section 3.3.4.1) and the diffusion equation proposed earlier97,98
were considered. The rates of surface oxygen exchange and spill-
over, intrinsic diffusion coefficients for surface and bulk oxygen and
respective activation energies, the overall quantity of oxygen-con-
taining species accumulated in the bulk, and characteristic times of
oxygen diffusion were estimated. It was found that oxygen spillover
between Pt and the support originates the fast oxygen exchange
observed over the supported Pt solid. The characteristic time of dif-
fusion of the oxygen localized in the subsurface layers was found to
be about 1 s with the oxygen in up to two monolayers. On the other
hand, the characteristic time of oxygen diffusion within the bulk was
about 20 s with up to 3–4 monolayers of oxygen accumulated in the
bulk.

3.3.5 Is OSC in three-way catalysts an equilibrium-controlled process?


The important question as to whether the chemical steps involved in
the transient oxygen storage and release processes previously
described (Sections 3.3.1 and 3.3.2) should be considered as
equilibrium controlled processes (reversible elementary steps) has
recently been addressed.23 It was found that reduction of ceria by CO
or H2 to form CO2 or H2O, respectively, is very important, and if not
included in proposed reaction schemes, a number of effects experi-
mentally observed under the dynamic conditions of OSC measure-
ment cannot be understood. For example:23 (i) the observed OSC is
lower in the presence of H2O and CO2 in the CO/He and O2/He
alternate pulse or step-change gas compositions; (ii) the observed
OSC depends on the CO/H2 concentration; (iii) when the catalyst is
exposed to a rich-lean step, the outlet gas composition remains rich
for some time after the inlet has switched to lean; (iv) when a previ-
ously oxidized catalyst is subjected to a short rich pulse followed by
stoichiometric operation, the catalyst emits a delayed pulse of CO/
H2 during the stoichiometric phase; (v) when a catalyst is operated
with an oscillating air/fuel ratio at slightly rich conditions, the

b1469_Ch-03.indd 168 4/8/2013 12:29:05 PM


b1469 Catalysis by Ceria and Related Materials

Investigation of the Oxygen Storage and Release Kinetics of TWCs 169

oscillations are absorbed by the OSC. The equilibrium-based kinetic


model also estimates the individual forward and backward rates of
each chemical step of the OSC, which can be used to quantify the
degree to which each chemical reaction step reaches real equilib-
rium within the few seconds of dynamic OSC behavior. Further
research on modeling DOSC with practical three-way catalyst com-
positions, where equilibrium-based kinetic models should be
included, is of paramount importance for the development of new
and more fundamentally based λ control and catalyst diagnosis
algorithms.23

3.4 Case Studies


3.4.1 Deterioration of oxygen storage and release kinetics
in CeO2 and CexZr1−xO2−δ due to chemical poisoning
The thrust of current research into automotive exhaust-gas purifica-
tion is the development of advanced TWCs with chemical formula-
tions to give tolerance of poisons and thermal durability, and at the
same time prolong improved catalytic performance. Under normal
working conditions, catalytic converters progressively undergo struc-
tural and morphological changes caused by thermal and chemical
deactivation, the main reasons for TWC deactivation, leading to the
loss of activity and selectivity.61,68–73,99–111
Chemical deactivation involves stopping reactants from reach-
ing active catalytic sites, and chemical modification of the sites due
to the binding of impurity elements present in the exhaust gas. The
latter include: (i) Pb, S, and Si from fuel additives; (ii) P, Ca, Zn, and
Mg from lubricant oil additives; and (iii) Fe, Cu, Ni, Cr, and Cd from
wear of the engine and exhaust tailpipe. The most important deac-
tivating effects have been ascribed to the accumulation of P, Pb, and
S.61,68–71,73,99–111 Nevertheless, the reduction of sulfur levels in gasoline
and the current use of unleaded gasoline imposed as control meas-
ures in fuel reformulation, have led to the partial mitigation of the
problem. P, Ca, and Zn have been detected in large concentrations
on the surface of aged TWCs.99–103 The presence of these elements

b1469_Ch-03.indd 169 4/8/2013 12:29:05 PM


b1469 Catalysis by Ceria and Related Materials

170 A. M. Efstathiou and S. Y. Christou

in TWCs is attributed to the hydrolytic degradation of the antiwear/


antioxidant zinc dialkyldithiophosphate (ZDDP) and other deter-
gent additives used in engine oil. Their presence in TWCs appears
as an impervious glassy Zn phosphate and an amorphous glaze-like
Ca phosphate overlayer with structure (Zn,Ca,Mg)3(PO4)2, CePO4,
Ce(PO3)3, or AlPO4.99,105–110,112–114 The intrinsic effects caused by the
presence of these compounds on the multiple functions of a TWC
are not yet well understood. The complexity of the chemical and
structural formulation of TWCs, and the variety of contaminant spe-
cies accumulated on or reacting with the components of a TWC,
make any correlation of their presence with the degree of TWC
deactivation a difficult task. It is important, therefore, to acquire
fundamental information on the mechanisms through which chemi-
cal deactivation of commercial TWCs occurs.
Larese and co-workers61,68,69,112,113 and Granados et al.70 reported
that the presence of CePO4 detected in vehicle-aged commercial
TWCs correlates with the deterioration of its OSR properties and
catalytic activity. The recovery of OSC and the catalytic properties
of aged commercial TWCs after CePO4 decomposition with increas-
ing calcination temperature in the 500–900 °C range was shown.68
It was found that the OSC and OSCC measured by the CO/O2
pulse-injection technique on an aged TWC (29,900 km) were two
times lower compared to those measured over a fresh (0 km) TWC.
The rates of oxygen storage and back-spillover occurring through
the participation of Ce(–Zr) oxides and noble metals were largely
and irreversibly suppressed. This was explained because a fraction
of the cerium cations was locked as Ce(III) in the form of CePO4,
which is unable to participate in the Ce(IV)/Ce(III) redox cycle.
CePO4 was also found to be very stable against calcination at 900 °C,
and proved to be significant in the chemical deactivation of a
TWC through severe deterioration of the OSC and TWC
performance.68
Xu et al.108 studied the poisoning effect of phosphorus in both
model and monolithic catalysts aged on bench-engine dynamome-
ters and vehicles. They found that the direct reaction of ceria with
P2O5 present in the exhaust gas or with other P compounds

b1469_Ch-03.indd 170 4/8/2013 12:29:05 PM


b1469 Catalysis by Ceria and Related Materials

Investigation of the Oxygen Storage and Release Kinetics of TWCs 171

deposited on the catalyst, such as Al and Zn phosphates, irreversibly


formed monazite cerium orthophosphate (CePO4). The formation
of the latter was facilitated under rich oxygen exhaust-gas condi-
tions, presumably due to Ce reduction to the 3+ state, and also to
the exposure of the TWC to temperatures higher than 800 °C. In
order to determine the level of deactivation caused by thermal and
chemical effects (e.g., the presence of P in CePO4), OSC data were
obtained in the 250–700 °C range over a 2 wt% Pd/CeO2 model
catalyst. The amount of stored oxygen decreased by 16–25% after
simultaneous thermal and P treatments compared to that obtained
for the thermally aged catalyst alone. This was ascribed to the forma-
tion of cerium phosphate in agreement with previous studies.114
Based on the previous findings, a reasonable approach to reveal
the mechanisms of TWC deactivation by chemical poisoning is
through the intentional deposition of contaminants under con-
trolled conditions on catalytic systems consisting of simpler compo-
sitions than those encountered in commercial TWCs. For this
reason, Granados et al.70 prepared phosphated CeO2 with P/Ce
ratios of 0.01, 0.02, 0.05, 0.1, and 0.2 after impregnation of CeO2
with appropriate amounts of (NH4)2HPO4 solution followed by cal-
cination at 600 °C. The aim of that work was to explore the chemical
nature of the P species in TWCs with respect to their negative effects
on the OSC. In the samples with P/Ce < 0.03, isolated orthophos-
phate species were found to be formed on the surface and in the
subsurface regions of the solids. There was a continuous decrease in
the OSC measured by H2/O2 pulses with increasing P/Ce ratio (0.1
to 0.3) compared to pure ceria (Fig. 3.8(a)). In the solid samples
with P/Ce > 0.03, crystals of CePO4 monazite were found to be
formed. OSC measurements indicated that once CePO4 predomi-
nates over the oxide surface, the rate of OSC depletion with increas-
ing P/Ce ratio becomes small, indicating that as soon as CePO4
crystals are formed, further incorporation of P does not result in a
deterioration of the OSC deeper in the bulk of the CeO2 crystals.70
TPIE results showed a significant suppression in the maximum
rate of 18O exchange with surface 16O atoms, and in the amount of
exchangeable surface and bulk oxygen with increasing P/Ce ratio in

b1469_Ch-03.indd 171 4/8/2013 12:29:06 PM


b1469 Catalysis by Ceria and Related Materials

172 A. M. Efstathiou and S. Y. Christou

Figure 3.8 (a) OSC (µmol O·gcat−1) measured over Ce, CeP0.01, CeP0.02,
CeP0.05, CeP0.1, and CeP0.2 solids by H2/O2 pulse experiments in the 400–600 °C
range as a function of the experimental P/Ce ratio determined by total reflection
X-ray fluorescence (TXRF). (b) Number of equivalent monolayers of atomic oxy-
gen species exchanged during the TPIE experiment (Ne, Eq. (3.26)) for the inves-
tigated solids. Adapted with permission from Granados et al.70 Copyright 2006
Elsevier.

the phosphated CeO2 solid. Moreover, the number of oxygen mon-


olayers exchanged was found to decrease continuously with increas-
ing P/Ce ratio in the phosphated samples (P/Ce ≤ 0.3, Fig. 3.8(b)).
However, for P/Ce > 3, Ne was similar regardless of the amount of P

b1469_Ch-03.indd 172 4/8/2013 12:29:06 PM


b1469 Catalysis by Ceria and Related Materials

Investigation of the Oxygen Storage and Release Kinetics of TWCs 173

incorporated in the ceria. This was due to the formation of larger


monazite crystals on the surface at the expense of isolated orthophos-
phate species, leaving part of the ceria surface uncovered, thus avail-
able for surface and bulk oxygen exchange, in agreement with the
OSC measurements obtained. After calcination of the phosphated
samples at 1000 °C, the OSR properties of all the P/CeO2 samples
deteriorated to a greater extent compared to samples calcined at
600 °C. This was related to the dramatic collapse of the BET area
and the texture of ceria, and to the formation of CePO4 crystals,
which predominated at the surface of the solid irrespective of P
loading.115
The role of P in the chemical deactivation of TWCs was also
investigated after deliberately poisoning CeO2 and Ce0.8Zr0.2O2 with
P with an atom ratio of P/(Ce+Zr) = 0.2.61 The alteration of the
redox chemistry of the Ce-based oxides containing P, relevant to the
OSR processes, was studied by means of TPIE and TIIE (18O/16O
isotopic exchange) techniques. It was found that not only the con-
centration but also the diffusion rates of surface labile and bulk
oxygen were significantly reduced after P poisoning due to the for-
mation of CePO4 at the surface/subsurface region of the solids. In
particular, TPIE studies revealed a substantial depletion in the con-
centration and equivalent number of monolayers of exchangeable
oxygen (Ne, Eq. (3.26)) after P poisoning. More precisely, the value
of Ne = 0.9 (T = 500 °C) dropped to 0.4 and 0.2 for CeO2 and
Ce0.8Zr0.2O2, respectively, after P addition, demonstrating a large
reduction in the rate of surface oxygen diffusion. At T > 600 °C,
several oxygen monolayers were found to be exchanged, indicating
that beyond this temperature diffusion of oxygen from the inner
bulk of the solid is the predominant process for supplying oxygen to
the surface.
TIIE experiments performed at 600 °C, Fig. 3.9(a), showed
that the ratio of the bulk diffusion coefficients (Dbeff, Eq. (3.29))
estimated for the uncontaminated and P-contaminated solids
(Cαo was considered to be the same) was found to be 4.0 and 2.8
for CeO2 and Ce0.8Zr0.2O2, respectively, suggesting that the spe-
cific rate of bulk oxygen diffusion is remarkably suppressed after

b1469_Ch-03.indd 173 4/8/2013 12:29:06 PM


b1469 Catalysis by Ceria and Related Materials

174 A. M. Efstathiou and S. Y. Christou

(a)

(b)

Figure 3.9 (a) TIIE experiment performed at 600 °C over Ce and CeP samples
according to the gas delivery sequence: 3% 16O2/He (600 °C, 2 h) → 3% 18O2/He
(600 °C, t). (Adapted from Larese et al.61) (b) Transient response curves of 16O2(g)
species during a TPIE experiment over CeZrc600, CeZrPc600, CeZrc1000, and
CeZrPc1000 solids. Adapted with permission from Larese et al.69 Copyright 2005
Elsevier.

P incorporation.61 An increase of calcination temperature from


600 to 1000 °C resulted in an improvement of the specific surface
and bulk oxygen diffusion rates in the phosphated Ce–Zr–O solid
(Fig. 3.9(b)). Calcination of the solid in air at 1000 °C caused

b1469_Ch-03.indd 174 4/8/2013 12:29:06 PM


b1469 Catalysis by Ceria and Related Materials

Investigation of the Oxygen Storage and Release Kinetics of TWCs 175

intense sintering and rearrangement of the CePO4 crystallites


with a fractional uncovering of the Ce0.8Zr0.2O2 particles from
CePO4. This resulted in the partial rejuvenation of the OSC of the
phosphated Ce0.8Zr0.2O2 solid. This effect was not observed for
CeO2, implying that CexZr1−xO2−δ solid solutions are excellent oxy-
gen storage materials for commercial TWC applications, display-
ing enhanced resistance against the P deterioration of their OSR
properties compared to pure CeO2.69
Experimental evidence61,68–70 for the negative effects of P deposi-
tion in Ce(–Zr) metal oxides on their redox properties strongly sup-
ports the fact that CePO4 in patches or islands surrounding the
Ce(–Zr)–O solid crystal particles constitutes one of the main reasons
for chemical deactivation in real vehicle-aged TWCs. CePO4 can
deteriorate the OSR properties of CeO2-based solids through the
formation of a crust of CePO4 covering the surface of the Ce oxides,
thus acting as the main barrier for oxygen storage and release
between the gas phase and the surface/subsurface of the Ce oxides.
Once CePO4 has formed, some of the surface Ce is locked out of the
Ce(IV)/Ce(III) redox process due to the strong P–O–Ce bonding.
Consequently, the presence of very stable Ce(III) in CePO4 results
in the deterioration of the OSR properties of ceria and CexZr1−xO2
since the rapid Ce(IV)/Ce(III) redox couple is a prerequisite for
the correct functioning of TWC. It is thus clear that the oxygen stor-
age capacity and oxygen mobility evaluated through dynamic
pulsed-/stepped-feed and 18O/16O isotopic exchange experiments
are valuable kinetic properties for assessing the deactivation of aged
TWCs.
To study the effect of Ca on the OSC, which is considered to be
a TWC contaminant, Granados et al.116 impregnated CeZrO4 mixed
metal oxide with a Ca nitrate solution to achieve nominal Ca/
(Ce+Zr) atom ratios of 0.01, 0.02, 0.05, 0.1, 0.2, and 0.5, followed
by drying and calcination at 600 °C. The OSCC was measured at
600 °C following the gas sequence 10% H2/Ar (1 h) → Ar → 1%
O2/Ar. Fig. 3.10(a) shows the OSCC values (squares) for the pre-
pared samples as a function of CaO molar fraction. The dashed
line represents the values expected if the CeZrO4 and CaO solids

b1469_Ch-03.indd 175 4/8/2013 12:29:06 PM


b1469 Catalysis by Ceria and Related Materials

176 A. M. Efstathiou and S. Y. Christou

(a) (b)

Figure 3.10 (a) OSCC values measured at 600 °C over CaO/CeZrO4 samples as a
function of CaO molar fraction (values in brackets correspond to the nominal Ca/
(Ce+Zr) atom ratio). (b) FTIR spectral region for which O−2 is visible (1126 cm−1)
as a function of Ca/(Ce+Zr) atom ratio. Adapted with permission from Granados
et al.116 Copyright 2008 Elsevier.

behave as a physical mixture with no interaction between them.


The OSCC would then be expected to be a linear combination of
the OSCC of pure CeZrO4 and CaO (2081 and 442 µmol O.g−1,
respectively). As can be seen in Fig. 3.10(a), a clear deterioration
of the OSCC was observed for Ca/(Ce+Zr) > 0.01. The degree of
deactivation was found to increase (up to ∼55%) with increasing Ca
loading. Figure 3.10(b) shows FTIR results where a band at 1126 cm−1
assigned to superoxide species (O−2) was detected. This infrared
band was observed only in CeZrO4, CeZrCa0.01, and CeZrCa0.02
solids. The opposite was true for Ca/(Ce+Zr) > 0.02, indicating that
a significant inhibition towards the formation of O −2 species occurs,
in agreement with OSCC measurements.
Physicochemical characterization of the Ca-contaminated sam-
ples by thermogravimetric analysis (TGA), TXRF, X-ray diffraction
(XRD), BET, and X-ray photoelectron spectroscopy (XPS) tech-
niques indicated that calcium, for low Ca loadings, existed in very

b1469_Ch-03.indd 176 4/8/2013 12:29:06 PM


b1469 Catalysis by Ceria and Related Materials

Investigation of the Oxygen Storage and Release Kinetics of TWCs 177

well-dispersed CaO domains of a few atoms thick.116 As the Ca load-


ing increased, larger fractions of the CeZrO4 surface were covered
by CaO. Once the amount of Ca required to totally cover the surface
was reached, further addition of Ca resulted in the formation of
large three-dimensional CaO crystals. It was hence proposed that
the main mechanism for the deterioration of the OSR properties of
CeZrO4 due to Ca poisoning involves surface fouling, irrespective of
the chemical nature of the CaO. The presence of the CaO physically
prevents contact between the O2 gas and the active surface sites of
the CeZrO4 during the oxygen storage process. It is also likely that a
solid-state reaction occurred to a limited extent between the two
phases, with a CaZrO3-like amorphous oxide forming at the grain
boundaries. This oxide can also chemically contribute to TWC deac-
tivation due to the withdrawal of Zr4+ cations from the CeZrO4 lat-
tice, which results in the formation of CexZr1−xO2−δ with worse OSR
properties.116
As previously mentioned, P, Ca, and Zn (lubricant oil additives)
are known to be the most concentrated contaminants in aged
TWCs.99–103,116 It is therefore necessary to explore not only their indi-
vidual effects but also their combined effect on the deactivation
mechanisms of TWCs. In a recent study,73 the effects of P along with
Ca or Zn deposited on a CexZr1−xO2 solid solution on the kinetics of
oxygen storage and release of the derived materials were investigated
by H2/O2 pulse-injection and TPIE techniques. Chemical poisoning
of the Ce0.5Zr0.5O2 solid by P (4.5 wt%), Ca (1.0 wt%), and Zn
(1.0 wt%) was found to result in a drastic decrease (up to 70%) in
both OSC (Fig. 3.11(a)) and OSCC (Fig. 3.11(b)), and to severely
affect the surface and bulk oxygen diffusion kinetics (Fig. 3.12). This
was attributed to the formation of CePO4 in the surface/subsurface
region of the solid, as well as to the formation of Ca3(PO4)2 and ZnO
on the surface of the solid, physically clogging small pores (< 10 nm)
as indicated by the significant loss in the BET (48–65%) and specific
pores volume (30–40%) of the solid. Another reason for the severe
reduction in the OSC and OSCC of P/Ce0.5Zr0.5O2 solid is the rear-
rangement of its surface composition resulting in a lowering of the
Ce/Zr surface ratio (XPS studies), and thus the concentration of

b1469_Ch-03.indd 177 4/8/2013 12:29:06 PM


b1469 Catalysis by Ceria and Related Materials

178 A. M. Efstathiou and S. Y. Christou

1000 Ce0.5Zr0.5O2 Ce0.5Zr0.5O2


P/Ce0.5Zr0.5O2
2000 P/Ce0.5Zr0.5O2
800

OSCC (µmol O g-1)


P-Ca/Ce0.5Zr0.5O2 P-Ca/Ce0.5Zr0.5O2
OSC (µmol Og-1)

1600
P-Zn/Ce0.5Zr0.5O2 P-Zn/Ce0.5Zr0.5O2
600
1200
400 800

200 400

0 0
550 650 750 850 550 650 750 850
Temperature ( oC) Temperature (oC)
(a) (b)

Figure 3.11 (a) OSC and (b) OSCC (µmol O·g−1) measured over uncontaminated
Ce0.5Zr0.5O2 solids and Ce0.5Zr0.5O2 solids contaminated with P, P–Zn, or P–Ca in the
550–850 °C range. Adapted with permission from Christou et al.73 Copyright 2011
Elsevier.

4.0 4.0
Ce0.5Zr0.5O2 Ce0.5Zr0.5O2
P/Ce0.5Zr0.5O2 730 T > 850oC P/Ce0.5Zr0.5O2
635 760 830
3.0 P-Ca/Ce0.5Zr0.5O2 3.0 P-Ca/Ce0.5Zr0.5O2
O18 O (mol%)
O2 (mol%)

725 780
P-Zn/Ce0.5Zr0.5O2 P-Zn/Ce0.5Zr0.5O2

2.0 2.0
16

16

1.0 1.0

0.0 0.0
400 500 600 700 800 T = 850oC 400 500 600 700 800 T=850 oC
Temperature (oC) Temperature ( oC)
(a) (b)

Figure 3.12 Transient response curves for (a) 16O2 and (b) 18O16O gaseous oxygen
isotopic species obtained during TPIE over fresh Ce0.5Zr0.5O2 solids and Ce0.5Zr0.5O2
solids contaminated with P, P–Zn, or P–Ca. Adapted with permission from Christou
et al.73 Copyright 2011 Elsevier.

labile oxygen species associated with the Ce–O–Zr ionic bonding.


X-ray photoelectron spectroscopy results showed that the incorpora-
tion of P within the subsurface region of Ce0.5Zr0.5O2 crystals and its
reaction with the solid, which results in the formation of CePO4,
caused the significant increase in Ce3+ (54–63% compared to 18% in
Ce0.5Zr0.5O2), which does not easily become re-oxidized.73
Figure 3.12 shows the transient response curves of 16O2(g) and
18 16
O O(g), obtained during a TPIE experiment over uncontaminated

b1469_Ch-03.indd 178 4/8/2013 12:29:06 PM


b1469 Catalysis by Ceria and Related Materials

Investigation of the Oxygen Storage and Release Kinetics of TWCs 179

Exchange Rate (µmol·g-1·s-1)


6.0
Exchange Rate (µmol·g-1·s-1)

Ce 0.5Zr0.5O2 6.0
Ce 0.5Zr0.5O2
P/Ce 0.5Zr0.5O2 P/Ce 0.5Zr0.5O2
5.0 P-Ca/Ce 0.5Zr0.5O2 5.0
P-Ca/Ce 0.5Zr0.5O2
P-Zn/Ce 0.5Zr0.5O2 P-Zn/Ce 0.5Zr0.5O2
4.0 4.0

3.0 3.0

2.0 2.0

1.0 1.0

16O18O
2
16O

0.0 0.0
450 550 650 750 850 450 550 650 750 850
Temperature (oC) Temperature (oC)
(a) (b)

Figure 3.13 Specific rates of (a) 16O2 and (b) 18O16O gaseous dioxygen formation
obtained during TPIE over the uncontaminated Ce0.5Zr0.5O2 solid and Ce0.5Zr0.5O2
solids contaminated with P, P–Ca, and P–Zn. Adapted with permission from
Christou et al.73 Copyright 2011 Elsevier.

Ce0.5Zr0.5O2 solids and Ce0.5Zr0.5O2 solids contaminated with P, P–Ca,


and P–Zn. The exchange process begins essentially via multiple het-
eroexchange (Type III) with the characteristic production of 16O2(g),
and as the temperature is increased, it proceeds via both multiple
and simple (Type II) exchange mechanisms (Fig. 3.12). Moreover,
deposition of P, P–Ca, or P–Zn on the Ce0.5Zr0.5O2 solid caused a clear
shift to higher temperatures of the maximum production rate of 16O2
and 18O16O by more than 100 °C, which suggests an increase in the
activation energies of the surface and bulk oxygen diffusion.
The number of exchangeable oxygen species and equivalent
surface monolayers obtained during the TPIE experiments for the
contaminated solids were found to be significantly lower than those
corresponding to the poison-free Ce0.5Zr0.5O2 solid. Additionally,
the specific rates of 16O2 (Fig. 3.13(a)) and 18O16O (Fig. 3.13(b))
production estimated as a function of reaction temperature show
that in the absence of contaminants, oxygen mobility largely
increases with temperature in the 450–650 °C range. At 600 °C, the
rate of 16O2 formation over the phosphated sample was found to be
six times lower than that over the poison-free Ce0.5Zr0.5O2 solid,
while after P–Ca or P–Zn deposition the decrease in the exchange
rate was by a factor of 12 and 26, respectively. Considering that in
the low-temperature range of the TPIE profiles 16O2(g) formation

b1469_Ch-03.indd 179 4/8/2013 12:29:07 PM


b1469 Catalysis by Ceria and Related Materials

180 A. M. Efstathiou and S. Y. Christou

(Fig. 3.13(a)) is derived from the first surface layer (Ne < 1), it
becomes clear that the presence of contaminants strongly sup-
presses surface oxygen diffusion. In addition, the rate of 18O16O(g)
formation for T < 750 °C was found to be up to 37 times larger for
Ce0.5Zr0.5O2 than for the poisoned samples. It is thus postulated that
bulk oxygen diffusion within the crystal particles of Ce0.5Zr0.5O2 is
strongly affected by chemical poisoning.
In order to ascertain the level of influence of the chemical con-
taminants on the kinetics of the surface and bulk oxygen diffusion,
the apparent activation energy barriers for surface oxygen diffusion
Esapp (estimated for Ne < 1) and bulk oxygen diffusion Ebapp (esti-
mated for Ne > 1) were estimated. The incorporation of P within the
subsurface and bulk of Ce0.5Zr0.5O2 solid oxide particles was found
to significantly increase Ebapp from 126.5 to 169.6 kJ·mol−1, while the
concurrent deposition of Ca or Zn along with P was found to
strongly suppress surface oxygen diffusion. In fact, Esapp increased
from 112.7 to 135.9 (P–Ca) and 149.3 kJ·mol−1 (P–Zn). It was thus
shown that the introduction of P, P–Ca, and P–Zn into Ce0.5Zr0.5O2
induces the formation of strongly bound surface oxygen, whereas P
(PO 43−) is mainly responsible for restricting bulk oxygen mobility.73
Although the deterioration of the OSR properties of CeO2 and
CexZr1−xO2 by P, Ca, and Zn poisoning has been extensively investi-
gated,61,68–70,73,115,116 most of this effort was focused on the effects of
contaminants on the support materials of TWCs in the absence
of noble metals. Only a very few studies focused on the effects of
chemical poisons on the oxygen storage and catalytic properties
of model three-way catalysts. The study by Tabata et al.100 on model
Zn-, Fe-, and Pb-poisoned catalysts showed that only Pb caused a
considerable decrease in the activity and selectivity of NO reduction.
On studying the effect of various dopants used in Rh/CeO2 as OSC
promoters, it was observed that Pb significantly damaged the OSC.12
Based on the findings of another study,111 Pb was found to dissolve
in CeO2 after its deposition at the level of 0.4 and 1 wt% on a model
1 wt% Rh/CeO2 catalyst. Lead was also found to cause severe dete-
rioration of catalytic performance for C3H6 oxidation, whereas CO
and NO conversions were only slightly affected, likely due to CO and

b1469_Ch-03.indd 180 4/8/2013 12:29:08 PM


b1469 Catalysis by Ceria and Related Materials

Investigation of the Oxygen Storage and Release Kinetics of TWCs 181

C3H6 competitive oxidation;111 the latter species were converted at


higher temperatures, where bulk reduction of ceria is hindered due
to the absence of Pb.
Heo et al.117 investigated field-aged Pd-based TWCs and observed
the gradual deactivation of their OSC and CO oxidation perfor-
mance under oxygen-rich conditions with increasing catalyst age.
The overall deactivation mechanism of a Pd-based vehicle-aged
TWC is schematically illustrated in Fig. 3.14. At the initial stage,
there is sintering of the Pd noble-metal particles and the formation
of CeO2 from the CexZr1−xO2 phase upon exposure to high tempera-
tures with a phase transformation to Ce2(SO4)3 and CePO4 by P- and
S-poisoning, respectively. The gradual deactivation of the TWC with
mileage is induced by a continuous increase in the amount of P
deposited on the surface of the TWC, eventually forming stable
CePO4, the primary deactivation route of OSC in TWCs, and finally
by the encapsulation of Pd into the pore structure of the support,
which results in the weakening of the Pd–Ce interactions and a
decrease in the number of active catalytic sites.117
In order to investigate the chemical poisoning effect in commer-
cial TWCs, Christou et al.118 performed a fundamental study on the
effects of accumulated P, P–Ca, and P–Zn on the surface of a 1 wt%
Pd/Ce0.5Zr0.5O2 model TWC on its structural, textural, morphologi-
cal, OSR, and catalytic properties. The deposition of P, Ca, and Zn

Figure 3.14 Deactivation process of Pd-based TWCs containing CexZr1−xO2.


Adapted with permission from Heo et al.117 Copyright 2011 Elsevier.

b1469_Ch-03.indd 181 4/8/2013 12:29:08 PM


b1469 Catalysis by Ceria and Related Materials

182 A. M. Efstathiou and S. Y. Christou

contaminants on the surface of the catalyst by the wet-impregnation


method followed by calcination at 850 °C resulted in the formation
of new phases of CePO4, Ca3(PO4)2, and ZnO due to chemical-type
structural transformations. These were considerably destructive for
the catalyst’s textural properties and Pd dispersion, where Pd sur-
face sites were partially covered by the contaminants. A remarkable
drop by 270–600 µmol O·gcat−1 in the OSC and OSCC measured by
H2/O2 pulses in the 550–850 °C range occurred after contamination
of the fresh Pd/Ce0.5Zr0.5O2 solid with P, P–Ca, and P–Zn.118 An
increase of the calcination temperature from 500 to 850 °C resulted
in only 10% loss in the OSC, whereas in the presence of P, Ca, or Zn,
the loss in OSC was significantly higher (ca. 45%). The Pd local
electronic environment was found to not be affected by the pres-
ence of these contaminants.118 The most significant decrease in the
OSC of Pd/Ce0.5Zr0.5O2 was observed at the lowest reaction tempera-
ture of 550 °C and in the presence of P–Zn (∼40% loss).
Bedrane et al.26,27 showed that the activation of oxygen on ceria
or ceria–zirconia supported catalysts is fast, with the noble-metal
particles acting as portholes for the subsequent migration and stor-
age of oxygen on the support. Thus, at low reaction temperatures Pd
dispersion plays a very important role since the transient rates of H2
oxidation (occurring during the H2/O2 pulses) on PdO and that of
the back-spillover of oxygen from the mixed metal oxide to the Pd/
PdO interface were found to decrease with increasing Pd particle
size.119 This was ascribed to the weakening of the metal/support
interactions, which are considered crucial for promoting oxygen
mobility and catalyst reactivity.27,119 Thus, the partial covering of the
Pd particles by P, Ca, or Zn compounds formed after deliberately
poisoning the catalyst resulted in an observed significant deteriora-
tion of the OSR properties of the Pd/Ce0.5Zr0.5O2 solid. The lowest
number of Pd surface sites per gram of catalyst and the lowest value
of the BET surface area for P–Zn–Pd/Ce0.5Zr0.5O2 were in agree-
ment with the lowest OSC and OSCC values obtained at 550 °C.118
The OSC obtained over Pd/Ce0.5Zr0.5O2-aged solid was found to
be four times larger compared to Ce0.5Zr0.5O2 alone (882 vs. 223 mmol
O.g−1, T = 550 °C).73,118 In the presence of P, Ca, or Zn contaminants,

b1469_Ch-03.indd 182 4/8/2013 12:29:08 PM


b1469 Catalysis by Ceria and Related Materials

Investigation of the Oxygen Storage and Release Kinetics of TWCs 183

even though the OSC and OSCC were significantly lower compared
to the uncontaminated solids, these values were found to be 7–10
times higher in the contaminated Pd/Ce0.5Zr0.5O2 than the contami-
nated Ce0.5Zr0.5O2 solids. For example, the OSC of P/Ce0.5Zr0.5O2 at
550 °C was 68 µmol O.g−1 and that of P–Pd/Ce0.5Zr0.5O2 was 665
µmol O·g−1. As the reaction temperature increased, the differences
between the poisoned Ce0.5Zr0.5O2 and Pd/Ce0.5Zr0.5O2 solids were
less pronounced. Also, the percentage reduction in OSC and OSCC
at 550 °C after poisoning the Ce0.5Zr0.5O2 solid was about 50–70%73
compared to 16–40% for the aged Pd/Ce0.5Zr0.5O2 solids.118 Thus,
the presence of a noble metal on the surface of the CexZr1−xO2 sup-
port diminishes to a considerable extent the negative effects of P,
Ca, and Zn poisons on the OSR properties. This is likely to be the
result of the promotion of hydrogen dissociation and activation
steps and of the back-spillover of oxygen from Ce0.5Zr0.5O2 to the
oxygen vacant sites formed during reduction of PdO to Pd0. This is
in agreement with the XPS findings, which showed that after poison-
ing the proportion of Ce(IV) in CexZr1−xO2 after oxidation decreased
from 82 to 37–46%, whereas in the presence of Pd it decreased to
∼66%. This is a very important finding, showing the positive effect
of Pd in preventing the deactivation of OSC in a Pd/CexZr1−xO2 cata-
lyst (Ce3+ is mainly associated with CePO4 formation), which was
confirmed by OSC measurements. Kroger et al.109,110 studied the
individual and combined effects of P and Ca on the OSC and cata-
lytic activity of supported Rh catalysts. A strong catalyst deactivating
effect was observed in the presence of P, whereas the addition of
P–Ca had a regenerating effect since the poisoning effect of phos-
phorus alone was reduced.
The role of an Fe contaminant in the chemical deactivation of a
1 wt% Pd–Rh/20 wt% CeO2–Al2O3 (PRCA) model TWC has been
investigated.71 H2/O2 pulsed-feed OSC measurements in the 500–
850 °C range indicated that OSC and OSCC (Fig. 3.15(a)) increased
by up to 40% in the presence of 0.1–0.3 wt% Fe. The amount
of oxygen stored in pure Fe2O3 corresponding to 0.3 wt% Fe was
found to be 80 µmol O·g−1. This amount cannot explain the OSCC
found in the 0.3 wt% Fe/PRCA solid measured at 700 °C, which was

b1469_Ch-03.indd 183 4/8/2013 12:29:08 PM


b1469 Catalysis by Ceria and Related Materials

184 A. M. Efstathiou and S. Y. Christou

Area band at 1126 cm-1


T=25 C

o
T=125 C

-
O O
16
16
o
T=225 C

e/ A

e/ A

A
2F CA

2F CA

A
C

C
0. PR

PR

0. PR

PR
PR

0. PR

PR
e/

e/
4F

4F
0.
(a) (b)

Figure 3.15 (a) OSCC (µmol O·gcat−1) measured by H2/O2 pulsed-feed experiments
over 0, 0.1, 0.2, 0.3, and 0.4 wt% Fe/PRCA solids in the 500–850 °C range. (b) Integral
band at 1126 cm−1 due to O−2 estimated after oxygen chemisorption at 25 °C followed
by TPD in Ar flow (125 and 225 °C) over PRCA, 0.2, and 0.4 wt% Fe/PRCA solids.
Adapted with permission from Lambrou and Efstathiou.71 Copyright 2006 Elsevier.

found to be equal to 345 compared to 208 mmol O·g−1 for the PRCA
catalyst; an increase in the OSCC of 27% was estimated after sub-
tracting the Fe2O3 contribution. This synergistic effect was investi-
gated by X-ray photoelectron spectroscopy, where it was found that
iron deposited on the Pd/Rh and CeO2 surfaces induced the devel-
opment of electronic interactions that influenced the oxidation
states of the noble metals and also enhanced the oxygen chemisorp-
tion on ceria, likely affecting to some extent the Ce–O–Ce bond
strength. It was concluded that Fe can act as an additional OSC
promoter under oxidizing conditions through the process: Fe →
FeO → Fe3O4 (FeO·Fe2O3) → Fe2O3, and not as a chemical poison of
the OSR properties of the solid. On the other hand, when the
amount of Fe deposited increased to 0.4 wt%, the OSCC deterio-
rated but remained slightly greater than that measured over the
PRCA catalyst (Fig. 3.15(a)). In situ DRIFTS oxygen chemisorption
at 25 °C followed by temperature-programmed desorption in Ar
(Ar-TPD) showed that after 0.2 wt% Fe deposition, the formation of
superoxide species on CeO2 was strongly enhanced (Fig. 3.15(b)).
By increasing the temperature in Ar flow during TPD, superoxide
species were found to desorb completely at T > 250 °C. By further

b1469_Ch-03.indd 184 4/8/2013 12:29:08 PM


b1469 Catalysis by Ceria and Related Materials

Investigation of the Oxygen Storage and Release Kinetics of TWCs 185

increasing the Fe loading to 0.4 wt%, the chemical bonding between


O−2 and the ceria surface was found to weaken, resulting in a
decrease of its population, which correlates well with the OSC
results.71
In situ DRIFTS spectra were also recorded during TIIE experi-
ments at 25 °C (3% 16O2/Ar (30 min) → Ar → 3% 18O2/Ar (t)) as a
function of time in an 18O2/Ar gas flow.71 The oxygen exchange was
found to proceed mainly via a multiple exchange mechanism.
Adding 0.2 wt% Fe in PRCA considerably increased the concentra-
tion of 18O−2 species compared to that in PRCA alone, with a con-
comitant depletion of the presence of adsorbed 16O−2. In contrast,
adding 0.4 wt% Fe in PRCA resulted in a significant decrease in the
concentration of exchangeable 16O−2 adsorbed species compared to
that obtained over the 0.2 wt% Fe/PRCA solid. However, the popu-
lation of these species was still higher than that found in the PRCA
solid. It has been suggested71 that an optimum Fe loading exists that
enhances the overall rate of the oxygen multiple heteroexchange.
According to that study,71 the presence of Fe likely causes the
enhancement of oxygen diffusion on the surface of ceria and a
decrease in the bond strength between binuclear oxygen and the
ceria surface. Comparing the OSCC and TIIE results it might be
suggested that binuclear oxygen participates in the oxygen storage
and release processes in both the PRCA and Fe/PRCA solids.71
In a continuation of the work mentioned above,71 which investi-
gated the role of accumulated Fe on the surface of TWCs as a chemi-
cal poison, catalytic activity studies were conducted over a Pd–Rh/
CeO2–Al2O3 model TWC in the presence and absence of Fe.120 It was
found that when Fe is deposited on the model TWC up to a level of
0.4 wt%, its activity towards CO and C3H6 oxidation and reduction of
NO by H2 does not deteriorate. Instead, it was found that iron signifi-
cantly improved the T50 parameter. It was suggested that small Fe
clusters in contact with the noble-metal particles, due to the lower
work function of Fe compared to Pd and Rh, likely act as a source of
electron flow towards the noble metals. This results in the existence
of their reduced state to a considerable degree under CO and C3H6
oxidation conditions, and also during reduction of NO by H2. Iron

b1469_Ch-03.indd 185 4/8/2013 12:29:08 PM


b1469 Catalysis by Ceria and Related Materials

186 A. M. Efstathiou and S. Y. Christou

seems to promote oxygen chemisorption via the modification of the


surface electronic properties of Pd and Rh, and likely via surface dif-
fusion of oxygen from the iron oxide clusters in contact with the
noble-metal particles. The latter significantly increased the activity of
the model TWC towards oxidation of CO and C3H6, and to a lesser
extent towards the reduction of NO by H2. The presence of Fe on the
surface of the model Pd/Rh-based TWC also created new active cata-
lytic sites for the reactions investigated.71 It was concluded that Fe
deposition on commercial TWCs, at least up to a level of 0.4 wt%, very
likely acts as a promoter rather than a poison of catalytic activity.120

3.4.2 Deterioration of the oxygen storage and release kinetics


in CeO2 and CexZr1-xO2-δ induced by thermal aging
The loss of surface catalytic sites of a three-way catalyst at high tem-
peratures (T > 900 °C) occurs via structural modifications of its vital
components (e.g., the noble metals and OSC materials) resulting in
the degradation of its activity and OSC performance. Noble-metal
sintering, which leads to a decrease in the number of active sites
(mol·gTWC−1), and the loss of the washcoat BET surface area, attrib-
uted to phase transformation and sintering of the CexZr1−xO2 and
alumina solid phases, are the usual structural modifications found in
TWCs.121,122 Sintering or microcrystal growth of both the noble metal
and the CexZr1−xO2 OSC materials at elevated temperatures results in
the loss of interfacial area between the OSC material and the Noble
Metal (NM), the latter promoting the oxygen storage and release
kinetics (synergetic effect).7,14,19 High-temperature aging of ceria
particles renders the bulk of the ceria inactive for the Ce4+ ↔ Ce3+
redox process. The driving force for sintering is the lowering of sur-
face energy, so that migration of atoms or crystallites leads to the
coalescing of small particles into larger ones, thus to crystallite
growth and the elimination of pores.102,123,124 Another mechanism for
noble-metal surface area reduction is the deep encapsulation of
metal particles by the support material, which is induced by the
collapse of the support’s pore structure and the migration of noble-
metal cations (e.g., Rh3+) into the alumina lattice.14,125–128 The main

b1469_Ch-03.indd 186 4/8/2013 12:29:08 PM


b1469 Catalysis by Ceria and Related Materials

Investigation of the Oxygen Storage and Release Kinetics of TWCs 187

parameters which affect the rate of sintering are: aging temperature


and gas atmosphere, metal nature and particle size (dispersion),
support textural properties, and the presence of promoters and
impurities. The rate of sintering increases with temperature, and it
is generally much higher in an oxidative than a reducing gas atmos-
phere. The presence of water vapor in the feed stream also acceler-
ates the crystallization and structural modification of metal-oxide
supports.103,123 It is now well understood that the development of
modern TWCs requires the production of a new generation of OSC
materials with high surface area and phase stability, and which can
endure high temperatures (1000–1100 °C) under aging gas environ-
ments, while maintaining very good oxygen storage characteristics.
Enhancing surface and bulk oxygen mobility, and improving the
thermal stability of ceria–zirconia-based OSC materials can be
achieved through the use of advanced preparation methods, com-
positional optimization, the incorporation of small amounts of
dopants (e.g., Y, La, Pr, or Nd) in their matrix, and the application
of appropriate redox treatments.13,14,19,48,49,54,72,121–156
The high-temperature degradation of the support materials
used in TWCs usually involves a phase transformation of γ-Al2O3 to
lower-area phases: δ-Al2O3 (900 °C) → θ-Al2O3 (1000 °C) → non-
porous α-Al2O3 (1200 °C).14,126 The formation of cerium aluminate
(CeAlO3) due to the reaction of Ce3+ with alumina under gas reduc-
ing conditions and at temperatures exceeding 900 °C has been also
observed in aged TWCs. This causes the deactivation of ceria-based
OSC materials because of the stabilization of Ce(III).19,111,127 The
formation of CeAlO3 was shown to be hindered when using Ce–
Zr–O mixed oxides in place of CeO2, and in the presence of a NM,
since its formation is retarded due to the intimate contact between
NM or zirconia with ceria.14,19,111,129,130 CexZr1−xO2 solids may also
undergo phase transformation and segregation into new Zr-rich
tetragonal phases (e.g., Ce0.20Zr0.80O2) or Ce-rich cubic phases (e.g.,
Ce0.82Zr0.18O2) at high temperatures, typically after prolonged expo-
sure at T > 1000 °C. XRD studies could be used to verify phase
demixing in CexZr1−xO2 because of the broadening and splitting in
the main diffraction peaks of the solid after aging.13,19,121,126,131 Partial

b1469_Ch-03.indd 187 4/8/2013 12:29:08 PM


b1469 Catalysis by Ceria and Related Materials

188 A. M. Efstathiou and S. Y. Christou

phase demixing of Ce0.50Zr0.50O2 after aging at 950 °C in an oxidizing


gas atmosphere into Ce0.82Zr0.18O2 and Ce0.20Zr0.80O2 was found to be
accelerated in the presence of a NM (Pt or Pd). It was found that Pd
enhanced the demixing process to a greater extent than Pt.132 With
Ce0.68Zr0.32O2, the NM was not observed to have any effect. This was
suggested to be due to metal/support interactions, which depend
on the crystal structure of the oxide.132
Christou et al.72 compared the OSC (Fig. 3.16(a)) and OSCC
(Fig. 3.16(b)) measured by H2/O2 pulses in the 450–750 °C range
for two CexZr1−xO2 solids (70 wt% ZrO2, 20 wt% CeO2, balanced
with La2O3 and Nd2O3) synthesized by two different commercial
proprietary co-precipitation methods (A and B), as a function of
calcination temperature, Tcalc (700–1100 °C) in air. A severe loss of
OSC and OSCC occurred with the increase of Tcalc from 700 °C or
850 °C to 1000 °C (Fig. 3.16), whereas an additional small decrease
was observed by further increasing Tcalc to 1100 °C. It was noticed
that as the reaction temperature increased, the percentage loss in
OSC and OSCC decreased. In particular, for CexZr1−xO2–A, a 92%
and 83% loss in OSC and OSCC, respectively, was observed at 450 °C
and Tcalc = 1000 °C, whereas a 40% and 35% loss in OSC and
OSCC, respectively, was observed at 750 °C. It is expected that in

600 800
CexZr 1-xO2-A-700 oC CexZr 1-xO2-A-700 oC
A-1000 oC 700 A-1000 oC
OSCC-H2 (µmol-O/g)

500
OSC-H2 (µmol-O/g)

A-1100 oC 600 A-1100 oC


400 CexZr 1-xO2-B-850 oC CexZr 1-xO2-B-850 oC
500
B-1000 oC B-1000 oC
300 B-1100 oC 400 B-1100 oC
300
200
200
100
100
0 0
450 550 650 750 450 550 650 750
Temperature ( oC) Temperature ( oC)
(a) (b)

Figure 3.16 Comparison of (a) OSC and (b) OSCC obtained in the 450–750 °C
range over CexZr1−xO2–A and CexZr1−xO2–B solids calcined at 700 °C (solid A),
850 °C (solid B), 1000 °C, and 1100 °C. Adapted with permission from Christou
et al.72 Copyright 2006 Springer.

b1469_Ch-03.indd 188 4/8/2013 12:29:08 PM


b1469 Catalysis by Ceria and Related Materials

Investigation of the Oxygen Storage and Release Kinetics of TWCs 189

the low-temperature range the rate-determining step for oxygen


storage and release is confined to the surface. Thus, the BET sur-
face area is expected to play an important role since the amount of
available surface oxygen per gram of solid is greater for high sur-
face area CexZr1−xO2 solids. On the other hand, at higher reaction
temperatures bulk oxygen diffusion towards the surface of the solid
becomes important and contributes to the transient rate of oxygen
release. CexZr1−xO2–B appears to be slightly better than CexZr1−xO2–A
for TOSC ≤ 650 °C. A greater loss in OSC and OSCC was observed for
CexZr1−xO2–A compared to CexZr1−xO2–B at Tcalc = 1000 °C. These
results are in harmony with the BET area measurements showing
greater loss in the BET area for CexZr1−xO2–A (Table 3.4). It is worth
mentioning that only a small phase segregation was observed for the
two solids at Tcalc = 1100 °C.
Figure 3.17(a) shows dimensionless (Z = 1 corresponds to
2 mol% O2) gas-phase transient isothermal response curves of 16O2,
18
O2, and 18O16O isotopic species obtained during TIIE at 700 °C
for a fresh CexZr1−xO2–A solid. The formation of 18O16O(g) starts
immediately after the switch 2%16O2/2%Ar/He → 2%18O2/He,
and passes through a maximum after about 15 min on stream.
Beyond this time, the rate of exchange of 18O from the gas phase
with 16O in the solid (surface/subsurface) decreases, and a tail
out to more than 80 min on stream is obtained. On the other
hand, the 16O2(g) transient response curve follows a smooth

Table 3.4 BET surface area and activation energies of surface oxygen mobility
(E sapp) and bulk oxygen diffusion (Eb) for CexZr1−xO2 solids.72
Solid BET (m2·g−1) Esapp (kJ·mol−1) Eb (kJ·mol−1)
CexZr1−xO2–A/700 °C 89 28 ± 2 104 ± 4
CexZr1−xO2–A/1000 °C 50 62 ± 3 105 ± 5
CexZr1−xO2–A/1100 °C 10 114 ± 4 116 ± 5
CexZr1−xO2–B/850 °C 90 44 ± 3 98 ± 3
CexZr1−xO2–B/1000 °C 56 47 ± 3 108 ± 5
CexZr1−xO2–B/1100 °C 16 112 ± 4 113 ± 5

b1469_Ch-03.indd 189 4/8/2013 12:29:09 PM


b1469 Catalysis by Ceria and Related Materials

190 A. M. Efstathiou and S. Y. Christou

2 mol% 18O2/He
1.0

Z 16 O2, 18 O2, 16 O18 O 0.8


T = 700oC
0.6 16
O2
18
O16O
0.4 18
O2

0.2

0.0
0 1000 2000 3000 4000 5000
Time (s)
(a)
2 mol%18 O2 /He
4.0
Rmax
3.5
CexZr1-xO2-A
(µmol/g.s)

3.0
550o C
2.5 700o C
2.0 800o C
O18O

1.5
RS.S.
16

1.0
R

0.5
0.0
0 1000 2000 3000 4000
Time (s)
(b)

Figure 3.17 (a) Dimensionless gas-phase transient isothermal response curves of


16
O2, 18O2, and 18O16O obtained during TIIE at 700 °C over a fresh CexZr1−xO2–A
sample. (b) Transient rate (µmol·g−1·s−1) of 18O16O production over a fresh
CexZr1−xO2–A sample during TIIE experiments at 550 °C, 700 °C, and 800 °C.
Adapted with permission from Christou et al.72 Copyright 2006 Springer.

exponential-like decay for about 40 min on stream. It should be


noted that the decay response curve of a non-adsorbing inert gas
(e.g., Ar) for the same switch lasts for about 2 min (not shown in
Fig. 3.17(a)). Thus, the observed 16O2(g) transient response curve
indicates that at the beginning of the experiment the exchange of

b1469_Ch-03.indd 190 4/8/2013 12:29:09 PM


b1469 Catalysis by Ceria and Related Materials

Investigation of the Oxygen Storage and Release Kinetics of TWCs 191

18
O with surface 16O takes place. Based on the material balance and
using the 16O2(g), 18O16O(g), and Ar response curves, it was esti-
mated that the amount of 16O exchanged was less than one equiva-
lent monolayer of surface oxygen. Therefore, up to the maximum
in the rate of 18O16O(g) formation, surface oxygen adsorption,
exchange, and diffusion processes prevail. On the other hand, the
reduced rate of 18O16O(g) formation following the peak maximum
shows a long tail that reaches a pseudo-steady state after about
83 min on stream (Fig. 3.17(a)). The latter originates from the
continuous but slow feed of the surface with 16O from the subsur-
face and bulk of the solid.
Figure 3.17(b) shows the evolution of the transient rate of
18 16
O O(g) formation obtained for CexZr1−xO2–A solid after applying
the TIIE technique at 550 °C, 700 °C, and 800 °C. In the time regime
where surface oxygen exchange and diffusion prevail, it is seen that
the rate of 18O16O(g) formation increases significantly with exchange
temperature. However, subsurface and bulk diffusion rates depend
to a smaller degree on reaction T (see transient rates at t > 3000 s,
Fig. 3.17(b)). R max and R s.s (pseudo-steady state exchange rate)
shown in Fig. 3.17(b) characterize the effect of Tcalc on the transient
rate and activation energies of the oxygen release kinetics from the
surface, and on the oxygen diffusion from the bulk to the surface
of the solids. It can be seen that there is a clear monotonic increase
of E sapp with increasing Tcalc (Table 3.4). For CexZr1−xO2–A, E sapp
increased from 28.0 to 114.0 kJ.mol−1 on increasing Tcalc from 700
to 1100 °C, while for CexZr1−xO2–B, E sapp increased from 44.0 to
112.0 kJ·mol−1. It was suggested that thermal aging in air caused a
drastic change in the surface morphology or surface composition
(e.g., La and Nd enrichment) of CexZr1−xO2-based solids, which
affected the energetics of surface oxygen mobility. The Rs.s. values
(see Fig. 3.17(b)) were used to estimate the activation energy (Eb)
of bulk oxygen diffusion (Table 3.4), which was found to only
slightly increase with increasing calcination temperature in the
range investigated.72
The aging effect on the OSC was also investigated over the Rh,
Pt, Pd, Ru, and Ir metals supported on CeO2 and Ce0.63Zr0.37O2,

b1469_Ch-03.indd 191 4/8/2013 12:29:09 PM


b1469 Catalysis by Ceria and Related Materials

192 A. M. Efstathiou and S. Y. Christou

where particular emphasis was given to the influence of metal parti-


cle size and the role of the metal/oxide interface.133 In almost all
metal supported catalysts investigated, a decrease in the OSC on
aging was found due to a decrease in the specific metal surface area
(significantly lower dispersion values were measured after aging),
although no direct correlation was found. As shown in Fig. 3.18(a),
the evolution of OSC as a function of temperature for both fresh
and aged Pt/CeO2 (aged at 800 °C in air) and Pt/Ce0.63Zr0.37O2
(aged at 1000 °C in air) follows the same trend. The OSC decreased
almost to the same extent over the whole temperature range. On
aging, the number and/or the accessibility of active sites involved in
the storage process decreased. The number of oxygen monolayers
involved in the storage process was calculated to be 5.2 and 4.2 for
the fresh and aged Pt/Ce0.63Zr0.37O2, respectively, while for the ceria-
supported catalysts oxygen storage was found to be restricted to the
surface (Ne = 0.6 and 0.5 for Pt/CeO2 in the fresh and aged states,
respectively).133 Similar observations were also made by Hori et al.124
regarding the effect of thermal aging at 1000 °C on the OSR proper-
ties of CeO2 and CexZr1−xO2 (x = 0.75, 0.50, 0.25) supported-Pt cata-
lysts. The CexZr1−xO2 sample aged at 1000 °C had an OSC more than
twice as high even when compared to the fresh ceria sample. Despite
the significant drop of 80% in the surface area of Ce0.75Zr0.25O2 after
aging, there was only a 30% drop in the amount of oxygen released.
Moreover, it was estimated that for the samples aged at 1000 °C the
amount of CO2 formed was about eight times more than the availa-
ble surface oxygen; a result that shows the participation of bulk
oxygen in CexZr1−xO2 particles.124 It is thus clear that even after
severe aging, bulk reduction of CexZr1−xO2 did occur. A comparison
of the redox behavior of CeO2 and CexZr1−xO2 solids reveals that at
moderate temperatures (T < 500 °C), the reduction of ceria is essen-
tially limited to the uppermost layer of the oxide. Thus, the deterio-
ration of textural properties of ceria after thermal aging dramatically
affects all the surface-related OSC properties. In Ce–Zr–O mixed
oxides, however, the redox behavior is strongly modified since
reduction occurs concurrently at the surface and in the bulk of the
solid. In this case, bulk oxygen is involved at a lower temperature

b1469_Ch-03.indd 192 4/8/2013 12:29:09 PM


b1469 Catalysis by Ceria and Related Materials

Investigation of the Oxygen Storage and Release Kinetics of TWCs 193

Figure 3.18 (a) Evolution of OSC as a function of temperature on Pt/CeO2 (‹)


and Pt/Ce0.63Zr0.37O2 (ƒ) solids: fresh catalysts = closed symbols, aged catalysts =
open symbols. (b) Aging effects on the OSC as a function of temperature for a
series of Rh/Ce0.63Zr0.37O2 solids: Rh/CeZr-1 (500 °C, air, a ), Rh/CeZr-2 (800 °C,
air, …), Rh/CeZr-3 (900 °C, air, ‹), Rh/CeZr-4 (1000 °C, H2, c). Adapted with per-
mission from Bedrane et al.133 Copyright 2002 Elsevier.

and compensates for the loss of exchangeable surface oxygen due


to the drop in surface area.
Figure 3.18(b) shows the aging effects on the OSC for a series of
Rh/Ce0.63Zr0.37O2 catalysts aged in air at 500 °C (Rh/CeZr-1), 800 °C
(Rh/CeZr-2), 900 °C (Rh/CeZr-3), and in H2 at 1000 °C (Rh/CeZr-
4). At low temperatures (200–300 °C), OSC is limited to the metal
particle and its very close vicinity, and depends on metal dispersion.
In particular, OSC was found to decrease with decreasing

b1469_Ch-03.indd 193 4/8/2013 12:29:09 PM


b1469 Catalysis by Ceria and Related Materials

194 A. M. Efstathiou and S. Y. Christou

metal dispersion: D(%)= 62, 55, 14, and 8 for samples 1, 2, 3, and 4,
respectively.133 The continuous OSC decrease with increasing calci-
nation temperature was ascribed to kinetic factors that control the
OSC process, with a higher aging temperature leading to larger NM
particles, larger ceria–zirconia grains, and lower rates for oxygen
diffusion and/or the creation of surface vacancies.128 Above 300 °C,
the OSC is due essentially to the participation of support. The unex-
pected enhancement of OSC for the fourth sample, when the oxide
starts participating in the OSC process, was ascribed to the severe
reduction of the solid in H2 during aging, which appears to be ben-
eficial to the overall reducibility of Ce0.63Zr0.37O2.133
To isolate the effects of Pt from those inherent in the OSC
material, Pt was deposited on pre-sintered CeO2 and Ce0.75Zr0.25O2
solids.134 The rate of CO2 production measured by CO/O2 step-
change OSC experiments on Pt/CeO2 was found to be linearly
dependent on Pt dispersion; however, this was not the case with Pt/
Ce0.75Zr0.25O2 for which the rate was independent of Pt dispersion
above a threshold of the support’s surface area. Although Pt was
found to sinter more readily on a Ce0.75Zr0.25O2 support, the disper-
sion effect was more than compensated by the enhanced availabil-
ity of bulk O2− within the mixed-oxide crystal particles. On a surface
area basis, the mixed-oxide support was 2.5 times more active com-
pared to ceria.134 The encapsulation of Pt metal particles by the
low-surface area CeO2–ZrO2–La2O3, besides Pt sintering, was the
reason why the OSC performance deteriorated after calcination of
the supported Pt catalyst at 900 °C.135 The latter resulted in a large
decline in the concentration of chemisorption sites, as shown by
temperature-programmed reduction in H2 (H2-TPR) and OSC
studies. At higher temperatures, the concentration of chemisorp-
tion sites on Pt particles, which relies heavily on the loading and
dispersion of Pt, was found not to be essentially responsible for the
enhancement of OSC, since the sample calcined at 700 °C had a
higher OSC compared to those calcined at 500 °C or 900 °C. This
shows that partial encapsulation of the metal by the support might
be beneficial for the OSC process. It was proposed that distortion
of the support’s lattice and other crystal defects observed at the

b1469_Ch-03.indd 194 4/8/2013 12:29:09 PM


b1469 Catalysis by Ceria and Related Materials

Investigation of the Oxygen Storage and Release Kinetics of TWCs 195

metal/oxide interface would indicate an interaction between the


metal and the mixed metal oxide, which is particularly relevant to
the reducibility and OSC behavior of the low-surface area support.
By creating crystal imperfections along the metal/oxide interface,
encapsulation of Pt enhanced the abundance of Ce3+ and oxygen
vacancies, not only on the surface but also in the bulk of the oxides,
thus increasing the concentration of labile oxygen and the rate of
oxygen release.135 Jen et al.128 also found a new peak in the low-
temperature range of 100–200 °C in the H2-TPR trace of Pd/CeO2–
ZrO2 after aging at 1050 °C, which was attributed to Pd-assisted bulk
reduction of ceria. This was probably induced by the incorporation of
Ce into the Pd microcrystals or vice versa, forming a Ce–Pd solid solu-
tion. High-resolution transmission electron microscopy (HRTEM)
and XRD studies conducted on Pt/CeO2 and Pd/CeO2 have shown
the occurrence of Ce–NM alloying phenomena after thermal
aging,136 whereas it appears that these may also exist in NM/CeO2–
ZrO2 systems.128,135
It was suggested that another mechanism through which aged
CexZr1−xO2 materials show improved redox properties compared to
ceria124 or to samples exposed to severe oxidation at 900 °C129 is the
segregation of a small fraction of Ce3+ from the mixed metal oxide
on reduction. This fraction subsequently formed a finely dispersed
ceria-rich phase decorating the surface of larger ZrO2 or mixed-
oxide crystallites on re-oxidation.124,129 Furthermore, in the presence
of NMs, besides structural electronic modifications induced by
strong metal/support interactions (SMSI) at the interface after
high-temperature reduction, these could also contribute to the
alteration of the redox properties of the solid, and, therefore, of the
OSC, in a way that oxygen activation and transfer are facilitated or
hindered in the metallic phase.49,133,134,136–138 Summarizing, the influ-
ence of thermal treatments on the activity and OSC performance of
NM/CexZr1−xO2 catalysts may be due to: (i) alloy formation between
NM and Ce, (ii) decoration or encapsulation of the NM by partially
reduced ceria, (iii) pure electronic interactions (SMSI), and (iv)
structural modifications of the mixed metal oxide (e.g., phase segre-
gation or transformation).49,124–138

b1469_Ch-03.indd 195 4/8/2013 12:29:09 PM


b1469 Catalysis by Ceria and Related Materials

196 A. M. Efstathiou and S. Y. Christou

One of the most challenging aspects in the behavior of


CexZr1−xO2 materials is their chemical and structural sensitivity to
thermal aging. Of particular interest are the reversible changes of
reducibility induced by appropriate modification of the thermal
aging conditions.13,15,48,139–152 The latter has been reported to be the
source of an unusual enhancement of the low-temperature reduci-
bility of sintered CeO2–ZrO2 and NM/CeO2–ZrO2 systems observed
after these were subjected to a two-step treatment: severe reduction
(SR) at T ≥ 900 °C followed by mild oxidation (MO) at T ≤ 600 °C.
The promotional effect of the SR/MO treatments was found to be
reversibly deactivated after a SR/SO (severe oxidation) sequence,
indicating the transient nature and the considerable stability of this
improvement in reduction after several SR/MO cycles.139,140 The
change in the redox response was attributed to structural modifica-
tions of the mixed metal oxide, generating highly mobile lattice
oxygen species.139–152 However, despite such favorable reduction
behavior, a strong deactivation of dynamic oxygen storage capacity
was observed, indicating the limitations of the H2-TPR method
employed for the investigation of OSC. Dynamic H2- or CO-OSC
measurements did not follow the same trend as the redox response
obtained from the H2-TPR technique over CexZr1−xO2 subjected to
various thermal treatments (e.g., SR/MO-SR/SO cycling). This
implies that the two techniques do not have a direct correspond-
ence. Dynamic OSC was found to strongly depend on sample pre-
treatment, the nature of the reducing gas (H2 or CO), and the
surface area.48,139,140 This underlines the critical influence that prepa-
ration and activation procedures have on the surface chemistry and
the final OSC behavior of Ce–Zr mixed metal oxides.
On studying the influence of an aging gas atmosphere on the
OSC of CexZr1−xO2 solids, Vidal et al.54,142,143 revealed that although
the samples resulting from the two aging treatments, SO and SR/
MO, had similar textural properties, their redox behavior was con-
siderably different. Aging in air at 900 °C (SO, 140 h) resulted in a
Ce0.68Zr0.32O2 solid with slow kinetics for both reduction and
re-oxidation, whereas SR/MO (950/550 °C) was found to increase
the rate of O2 uptake and improve the OSC at low temperatures by

b1469_Ch-03.indd 196 4/8/2013 12:29:09 PM


b1469 Catalysis by Ceria and Related Materials

Investigation of the Oxygen Storage and Release Kinetics of TWCs 197

Table 3.5 OSC (mL·g−1) of fresh Ce0.68Zr0.32O2


(HS), after oxidative aging (LS) and severe
reduction (SR) (1 h, 5 vol% H2/Ar) measured by
O2-pulse experiments at 427 °C.54

Tred (°C) HS LS SR

200 0.2 0.2 1.5


350 2.0 1.2 6.9
500 9.5 6.7 10.2
700 13.0 9.8 13.4
900 15.6 13.9 15.3
1000 16.1 16.1 16.5

up to 7.5 times (Table 3.5) compared to the fresh (HS) and air-
aged (LS) solids after decreasing the reduction temperature.54
The degree of promotion of the reducibility by redox cycling was
found to increase with increasing Zr content in the CexZr1−xO2
solid. However, the OSC was limited due to the small amount of Ce
present.142 Although thermal aging in air was found to negatively
influence H2-TPR traces of the LS samples compared to those of
HS ones, remarkably, the favorable effects of the redox aging on
the reduction behavior were more pronounced in the low-surface
area CexZr1−xO2 solids compared to the fresh high-surface area
ones.143
Raman spectroscopy can be used to provide insights into the
origin of these observations with respect to modifications in the
oxygen sublattice (Fig. 3.19). For the fresh Ce0.68Zr0.32O2 sample
(HS), it was observed that the F2g mode (473–478 cm−1), which is
characteristic of a perfect fluorite lattice, was broad suggesting a
partial breaking of the selection rules due to the small particle size.
The increase in the intensity of the F2g mode after oxidative aging
(LS) was consistent with sintering of the sample, which apparently
leads to a relatively ordered situation with an oxygen polyhedron
around the metal cations. However, for the sample subjected to SR,
the broadening of the band at 473 cm−1 cannot be due to the small
particle size effect since this was treated at 950 °C. It might be

b1469_Ch-03.indd 197 4/8/2013 12:29:09 PM


b1469 Catalysis by Ceria and Related Materials

198 A. M. Efstathiou and S. Y. Christou

Figure 3.19 Raman spectra of the Ce0.68Zr0.32O2 samples. Adapted with permission
from Vidal et al.54 Copyright 1999 Elsevier.

indicative of breaking the local symmetry around the metal cations,


either due to high structural disorder in the cubic oxygen sublattice
or to a different geometry for the Ce–O and Zr–O bonds. It was sug-
gested that such a phenomenon, which may generate loosely bound
mobile oxygen atoms in the lattice, could be responsible for the
improved redox properties of the SR sample. HRTEM studies also
showed the formation of a new phase in which some of the linear
unit-cell dimensions doubled in size. This might be caused by the
displacement of O ions from the fluorite positions, ordering of the
Ce and Zr cations, or both of these in concert.54
It has been well established that structural changes in CexZr1−xO2
induced by thermal aging are associated with modifications in the
redox properties, although the relation between aging conditions,
nanostructural properties, surface chemistry, and redox behavior is
rather complex. These structural changes mainly consist of the for-
mation or destruction of the ordered pyrochlore-related cationic
Ce–Zr sublattice during aging. Starting from fresh ceria–zirconia
samples, typically exhibiting a random distribution in their cationic
sublattice, in the intermediate compositional range of the equilib-
rium phase (CeO2–ZrO2 system), there exist only cubic CaF2-type
(CexZr1−x)O2 phases above 1550 °C. In the range x = 0.2–0.7, when

b1469_Ch-03.indd 198 4/8/2013 12:29:09 PM


b1469 Catalysis by Ceria and Related Materials

Investigation of the Oxygen Storage and Release Kinetics of TWCs 199

the cubic CaF2-type CexZr1−xO2 phase is cooled in a furnace, solid


solutions with a tetragonal t′-CeZrO4 form appearing as a metastable
single phase can be formed, maintaining the random distribution of
Ce and Zr ions through a cation-diffusionless phase transition with-
out phase separation. After high-temperature reduction of the
t′-CeZrO4 phase, Ce2Zr2O7 has a pyrochlore structure, which pos-
sesses an ordered arrangement of Ce and Zr cations. A mild oxida-
tion of the pyrochlore-type precursor at 600 °C results in the
formation of the κ-CeZrO4 phase, which maintains the ordered
arrangement and releases oxygen at low temperatures.
As revealed by Raman, X-ray absorption fine structure (XAFS),
XRD, HRTEM, and high-angle annular dark-field scanning trans-
mission electron microscopy (HAADF-STEM) studies, the presence
or absence of the pyrochlore-related κ-CeZrO4 phase in aged
CexZr1−xO2 solids is considered to play a key role in determining
their redox response. CeO2–ZrO2 samples exhibiting an ordered
arrangement of Ce and Zr cations (κ-like phase) have been found to
show enhanced reducibility, as opposed to samples with a disor-
dered distribution in their cationic sublattice and which have a sup-
pressed OSC. Thermal aging treatments applied to CexZr1−xO2 solids
subjected to SR (T ≥ 900 °C) and a subsequent MO (T ≤ 600 °C) have
induced a cation-ordered arrangement of Ce4+ and Zr4+, like the one
in pyrochlore-related structures. The opposite happens for SO (T ≥
900 °C), which leads to a random distribution of Ce4+/Zr4+in the cati-
onic sublattice.144–153
Extended X-ray absorption fine structure (EXAFS) and XRD
studies performed on CeO2–ZrO2 (molar ratio Ce : Zr = 1 : 1) sug-
gested that there is a relation between the OSC and the local struc-
ture.149 The atomically homogeneous κ-CeZrO4 solid solution phase
showing an ordered cation environment was found to have the
highest OSC among the investigated samples, even though its BET
area was extremely low (1 m2·g−1). After aging in an oxidative gas
atmosphere at T > 1000 °C, the Ce/Zr ordered cation arrangement
collapsed. In particular, the OSC of the sample aged at 1200 °C was
about 40% of that of the fresh sample, where phase separation was
noticed. The latter results indicate that the OSC is strongly

b1469_Ch-03.indd 199 4/8/2013 12:29:10 PM


b1469 Catalysis by Ceria and Related Materials

200 A. M. Efstathiou and S. Y. Christou

dependent on the atomic structure of the CeO2–ZrO2 mixed metal


oxide.149
HRTEM images have revealed changes in the morphology of a
fresh low-surface area Ce0.62Zr0.38O2 sample after SR/MO (950/500 °C)
treatments, and verified the existence of the κ-phase, which is char-
acterized by a spatial frequency (lattice spacing), as opposed to that
observed in the starting material (compare Figs. 3.20(a) and (c)).
This is indicative of the characteristic ordering in a Ce–Zr sublat-
tice.150,151,153 However, after a SR/SO (950/950 °C) treatment the
characteristic contrasts of the κ-phase were not found, indicating a
transition to a disordered Ce–Zr sublattice. Likewise, changes in the

Figure 3.20 HRTEM images of (a) fresh, (b) SR/MO-aged, and (c) SR/SO-aged
Ce0.62Zr0.38O2 samples, and (d) their respective H2-TPR response curves. Adapted
with permission from Yeste et al.150 Copyright 1999 Elsevier.

b1469_Ch-03.indd 200 4/8/2013 12:29:10 PM


b1469 Catalysis by Ceria and Related Materials

Investigation of the Oxygen Storage and Release Kinetics of TWCs 201

morphology of the aged solid particles were also observed with


rounded ill-defined facets in the SR/SO sample, Fig. 3.20(b), and a
much better defined external shape with well-developed (111) fac-
ets in the SR/MO sample (Fig. 3.20(c)). The nanostructure of the
two thermally aged Ce0.62Zr0.38O2 solids, specifically the amount and
the surface presence of the κ-like phase, was found to significantly
affect their redox and OSC properties. For Tred ≤ 500 °C, the
enhanced reducibility of the SR/MO-aged sample compared to the
SR/SO-aged sample (Fig. 3.20(d)) was proposed to be kinetically
controlled by its surface structure, mainly consisting of the κ-like
phase, which is responsible for the faster surface activation of H2 gas
and is the rate-controlling step of the overall reduction process, as
confirmed from using oxide-supported Rh catalysts. In contrast, for
Tred ≥ 700 °C the observed differences in the degree of reducibility
and reduction for the SR/MO-aged sample were higher than those
for the SR/SO-aged sample. These differences were interpreted as
due to thermodynamic factors correlated with the total amount of
κ-phase present in the aged sample.150,151,153 CO-OSC and H2-OSC
data showed that the SR/MO-aged sample exhibited a superior
reducibility than the SR/SO one in the 200–900 °C range.153
Based on the above, the redox behavior and OSC of ceria–zirconia
mixed metal oxides critically depend on the nanostructure and
compositional homogeneity, and more specifically on the order/
disorder exhibited by the cationic sublattice. The higher the com-
positional homogeneity corresponding to the ordered pyrochlore-
related κ-phase, the higher the reducibility of the mixed oxide.144–154
The latter might be because the enhancement of the homogeneity
of the Ce–Zr–O matrix in Ce4+ and Zr4+ cations could facilitate the
valence change of Ce4+ (0.97 Å) to Ce3+ (1.14 Å) during the oxygen
release process by compensating for the stress energy arising from
the volume expansion, which would restrict any further valence
change in Ce. In addition, because of the atomically homogeneous
introduction of Zr4+ into the cubic CeO2 lattice, the Ce–O bond
length is shortened, as indicated by a decrease in the lattice con-
stant of CexZr1−xO2−δ, and the configuration of the oxygen around
Zr4+, which has a more centrosymmetric 8-fold coordination. This

b1469_Ch-03.indd 201 4/8/2013 12:29:10 PM


b1469 Catalysis by Ceria and Related Materials

202 A. M. Efstathiou and S. Y. Christou

coordination is generally considered to be spatially tight and unsta-


ble, because the ionic radius of Zr4+ (0.84 Å) is much smaller than
that of O2− (1.38 Å). On the other hand, the oxygen environment
around the Ce is also considered to be tight, because the Ce–O
bond in CeZrO4 is much shorter than in CeO2 alone. This instability
of the oxygen environment around Ce4+ and Zr4+ in κ-CeZrO4 gener-
ates active oxygen species, which are responsible for an improved
OSC.13,149,155,156

3.4.3 Regeneration of the oxygen storage and release properties


of aged commercial and model three-way catalysts
Understanding a catalyst’s deactivation is of crucial importance for
TWC technology since a significant enhancement in three-way cata-
lyst durability is expected. Furthermore, it would lead to a suitable
TWC regeneration methodology. Spent TWCs are nowadays com-
pletely destroyed and treated as potential resources of platinum-
group metals (PGMs). A suitable regeneration method would
contribute to preserving the raw materials and minimizing waste
compared to the hydrometallurgical procedures used to recover
PGMs. Despite the exclusive use of three-way catalysts in gasoline-
driven cars since the mid-1970s, only a few attempts have so far been
made to regenerate the OSR properties and catalytic performance
of deactivated TWCs using economically viable and environmentally
friendly methodologies. These attempts can be classified into the
following categories: (i) methods that apply thermo-chemical treat-
ments using O2, H2, or Cl2 gases157–159 or other Cl-containing rea-
gents, e.g., HCl, CCl4, and 1,2-dichloropropane (DCP),159–161 and (ii)
methods that involve washing the catalyst with liquid solvents, such
as weak organic acids or other chelating agents.101,162–172
Oxygen-rich or hydrogen-rich gas mixtures and combinations of
them at elevated temperatures (e.g., 500–700 °C) have been used to
improve the dispersion of NMs supported on various metal-oxide
carriers. The mechanism through which the thermal/oxygen treat-
ments improve metal dispersion includes the formation and split-
ting of noble-metal oxides (MOx), which after reduction form

b1469_Ch-03.indd 202 4/8/2013 12:29:10 PM


b1469 Catalysis by Ceria and Related Materials

Investigation of the Oxygen Storage and Release Kinetics of TWCs 203

smaller metal crystallites.157,158 The addition of gaseous chlorine in


oxygen gas mixtures has also been found to reverse the effects of
thermal sintering on precious metals through provoking the forma-
tion of mobile intermediate species, e.g., halide complexes of the
MOxCly type (M: Pt or Pd).173 The latter species are volatile, leading
to redispersion and restructuring of the active metal particles by
breaking large metal agglomerated clusters (> 50 nm) into smaller
ones.157–161,173 The enhancement of noble-metal dispersion results in
an increase of the number of active metal sites and also of the inter-
facial contact with the CexZr1−xO2 OSC component, thus improving
oxygen storage capacity160,161 and the rates of the main TWC
reactions.157–161
Investigations by Birgersson et al.157–159 aimed at finding a
regeneration method for the catalytic activity of aged Pd–Rh com-
mercial TWCs towards CO, NOx, and HC conversions by applying
thermal treatments under oxidizing, reducing, and chlorine-rich
gas environments. The influence of gas composition (10–100
vol%), treatment temperature (300–700 °C), time-on-stream (30–
600 min), and catalyst properties (e.g., NM loading, degree of
sintering, and fouling) were investigated. H2-gas treatments proved
to be ineffective,158 while after oxygen-rich gas treatments (20 vol%
O2/N2), the NM dispersion and catalyst performance were found
to be improved.157,158 Moreover, it was found that these treatments
are more efficient on catalysts with high NM loading and which
had experienced large sintering. The removal of combustible con-
taminants from the catalyst surface and the small decrease in the
size of the large noble-metal agglomerates (∼100 nm), as evidenced
by transmission electron microscopy (TEM) and energy dispersive
X-ray spectroscopy (EDS) after O2-gas treatments at 500 °C and 700 °C,
might account for the small shift observed towards lower tempera-
tures in the light-off curves. No apparent improvements were seen
at lower treatment temperatures (e.g., 300 °C).157,158
The addition of chlorine to the oxygen gas atmosphere (0.8 vol%
Cl2/18 vol% O2/N2) facilitated the restructuring of the noble metals
Pd and Rh, and further improved the rate of NM redispersion even
for catalysts with less sintering and lower NM loading. TEM images

b1469_Ch-03.indd 203 4/8/2013 12:29:10 PM


b1469 Catalysis by Ceria and Related Materials

204 A. M. Efstathiou and S. Y. Christou

showed that after applying the oxy-chlorination gas treatment, the


largest noble-metal agglomerates (up to 100 nm) fractured or split
into smaller particles (20–70 nm).157,158 H2-TPR studies also revealed
indirectly the presence of new Pd and Rh particles on the surface,
of size 20–30 nm, in agreement with TEM studies.158 Metal redisper-
sion was considered to be the result of the formation of MOxCly
halide species, which are redistributed and re-adsorbed on the sup-
port surface and form stable complexes.158,173 To prevent the loss of
such volatile MOxCly intermediate species, the treatment tempera-
ture should be kept below 500 °C and dilute chlorine-containing gas
mixtures should be used.158 However, traces of chlorine bound to
the bulk of the washcoat were found to invoke a short induction
period in the catalyst’s activity during the first few hours on the reac-
tion stream, and a higher conversion rate could be obtained only
after prolonged exposure to the exhaust gas. It is known that the
presence of Cl may affect the behavior of ceria-containing catalysts
due to the formation of CeOCl, which stabilizes Ce as Ce3+ and
modifies its ability to transfer reductant molecules at the metal/
metal oxide interface.160 Nevertheless, the performance of an aged
catalyst towards the three major pollutants present in the exhaust
gas was found to improve significantly after the oxy-chlorination gas
treatment. The high steam content in the exhaust gas proved to be
an effective way of removing residual chlorine.157
The potential effects of applying the oxy-chlorination gas treat-
ment on the redox, OSR, and catalytic properties of bimetallic Pd–
Rh (9 : 1)160 and Pt–Rh (3 : 1)161 model catalysts supported on 20 wt%
Ce0.5Zr0.5O2/Al2O3 (CZA) were investigated. Prior to applying the
regeneration gas treatment, the fresh catalysts were aged at 700 °C in
air for 3 h. Thermal aging was found to cause a considerable loss
(63–90%) of metal dispersion. It is worth pointing out that the aging
conditions were chosen so as to cause negligible sintering to the sup-
port particles as confirmed by BET and XRD measurements. After
aging, the OSC and OSCC measured in the 400–500 °C range by H2/
O2 pulses had decreased (Fig. 3.21). After increasing the reaction
temperature to 700 °C and 900 °C, the fresh Pd–Rh (Fig. 3.21(a)) and
Pt–Rh (Fig. 3.21(b)) catalysts stabilized at 400 °C probably underwent

b1469_Ch-03.indd 204 4/8/2013 12:29:10 PM


b1469 Catalysis by Ceria and Related Materials

Investigation of the Oxygen Storage and Release Kinetics of TWCs 205

Figure 3.21 OSCC (µmol O·g−1) measured by H2/O2 pulses over fresh, aged
(A973), and oxy-chlorinated (A973Cl) samples of (a) Pd–Rh/CZA and (b) Pt–Rh/
CZA. Adapted with permission from Daley et al.160 and Anderson et al.161 Copyrights
2005160 and 2006161 Elsevier.

sintering due to their exposure at high temperatures, and the OSC


and OSCC values were found to be even lower than those observed
for the corresponding aged catalysts. The reduction in OSC is the
result of the loss of the interfacial contact area between the noble
metals and the oxygen storage components due to NM sintering.
The oxy-chlorination gas treatment comprised injecting di-
chloropropane (DCP) into an air flow during the first hour of an

b1469_Ch-03.indd 205 4/8/2013 12:29:10 PM


b1469 Catalysis by Ceria and Related Materials

206 A. M. Efstathiou and S. Y. Christou

isothermal heating of the sample at 550 °C for 2 h.160,161 The oxy-


chlorination gas treatment was able to reverse the negative effects of
thermal aging on metal dispersion for Pd–Rh/CZA160 and Pt–Rh/
CZA161 model TWCs, leading to a recovery of metal dispersion
greater than 100%. As a consequence, the OSC and OSCC of the
oxy-chlorinated samples were found to have increased remarkably
compared to those for the aged and fresh catalysts (Fig. 3.21). The
catalytic activity for the CO/O2 and NO/H2/O2 probe reactions was
shown to have improved significantly160,161 due to the increase in the
number of available active metal sites previously lost during aging
(metal sintering). As a consequence, interactions between the metal
particles and the Ce–Zr–O mixed metal oxide (the OSC material)
had increased.
For the Pt–Rh bimetallic system, the regeneration procedure
was applied to two model catalysts with different NM loadings,
ca. 0.6 and 1.0 wt%.161 Improved OSC, OSCC, and light-off behav-
ior were only apparent after applying the oxy-chlorination treat-
ment for the low-loading noble-metal catalysts. Differences between
the two loadings were probably due to the degree of chlorination
of Pt during treatment, and the different extent to which volatile
species interact with the alumina and OSC components. Selective
redispersion of smaller crystallites can be explained on the basis of
a preferred interaction of Cl with low-coordination metal sites,
which exist in smaller metal crystallites having a large proportion
of high-index-type facets with high concentrations of step, kink,
and edge sites. Highly coordinated metal atoms present in low-
index-type facets in large metal crystallites would be less likely to
undergo reactions with Cl under similar conditions. Moreover, any
improvements in metal dispersion are only beneficial if the redis-
persed metal particles are deposited in locations where intimate
contact with the CexZr1−xO2 support rather than with alumina is
achieved.160,161
These studies focused on inverting only thermal deactivation in
TWCs. Nevertheless, access of Cl to the washcoat components might
be hindered by the accumulated contaminants deposited on aged
TWCs. In this case, the redispersion process might be inhibited by

b1469_Ch-03.indd 206 4/8/2013 12:29:10 PM


b1469 Catalysis by Ceria and Related Materials

Investigation of the Oxygen Storage and Release Kinetics of TWCs 207

the overlayer of contaminants covering the washcoat surface, pre-


venting, therefore, an increase in the lifespan of an aged TWC.
The removal of the accumulated contaminants from the surface
of a TWC can be achieved through a second class of treatments,
such as the use of chemical extractants (e.g., acetic, citric, and oxalic
acids) or chelating agents (e.g., EDTA, NTA, and S,S-EDDS),
through dissolution of the contaminant species, or complexation of
their metal ions formed in solution. Most of the contaminant com-
pounds formed during TWC operation, e.g., sulfate and phosphate
salts, are readily dissolved in acidic environments; however, weak
acids are preferred so as to prevent destruction of the washcoat
material. Alternatively, complexing agents act as chelate ligands
having a high affinity to form stable metal-ligand complexes in solu-
tion. The regeneration of an aged TWC by washing is through
removal of contaminants, inhibitors, and fouling agents from its
surface, thus increasing the number of exposed surface catalytic
active sites.101,162–172 Kim et al.171 tested both thermal/gas and acid-
aqueous solution treatments for regenerating spent TWCs by remov-
ing volatile organic compounds (VOCs). It was found that the
regain in catalytic activity was highly dependent on the treatment
conditions. More specifically, the acid-aqueous treatments were very
useful for improving catalytic activity due to the removal of various
contaminants, whereas the thermal/gas treatments reduced the
activity of aged TWCs.171
The efficiency of weak organic acids (acetic, oxalic, citric, NTA,
EDTA, and S,S-EDDS) for improving the OSR and catalytic proper-
ties of severely aged commercial TWCs (83,000–242,000 km mile-
age) through the removal of various contaminants (such as P, S, Zn,
Ca, Fe, Cr, Mn, Pb, Ni, or Cu) has been systematically investi-
gated.164–172 The leaching conditions, e.g., washing time, liquid flow
rate, concentration, and temperature of the acid solution, were
optimized for acetic acid162 and EDTA,167 so as to give the best effi-
ciency for the removal of contaminants, while at the same time pre-
venting the destruction of the washcoat material. Oxalic acid, which
was the strongest among the acids examined166,167 and had the larg-
est dissolution efficiency for phosphate and sulfate salts, was indeed

b1469_Ch-03.indd 207 4/8/2013 12:29:10 PM


b1469 Catalysis by Ceria and Related Materials

208 A. M. Efstathiou and S. Y. Christou

Figure 3.22 (a) Removal of P (%) after washing the aged TWC (83,000 km) with
acetic, citric, NTA, EDTA, and oxalic acid solutions (0.1 M). Adapted with permis-
sion from Christou et al.165 Copyright 2006 American Chemical Society. (b) XRD
patterns for the aged and EDTA, citric, and oxalic acid treated TWC samples
(cordierite (•); CexZr1−xO2 (d); MgZn2(PO4)2 (…); CaZn2(PO4)2/AlPO4 ({);
CePO4 (V). Adapted with permission from Christou.169 Copyright 2006 Central
Library of the University of Cyprus.

the most efficient among the leaching solutions investigated in


extracting P- (Fig. 3.22(a)) and S-containing species, as revealed by
inductively coupled plasma–atomic emission spectroscopy (ICP-
AES) and XPS analyses.166,167 Citric acid, NTA, EDTA, and S,S-EDDS

b1469_Ch-03.indd 208 4/8/2013 12:29:10 PM


b1469 Catalysis by Ceria and Related Materials

Investigation of the Oxygen Storage and Release Kinetics of TWCs 209

appeared to be more efficient in removing Pb, Zn, Ca, Mn, Fe, Cr,
or Ni metal contaminants through complexation, while acetic acid
proved to be the least effective.164–167,170 Combinations of oxalic acid
with citric acid or EDTA aiming at effectively extracting a wider
range of contaminants, were found to be less efficient and led to
intermediate removal efficiencies than those obtained by any of the
individual acid washing solutions investigated.166,167,169
XPS analysis also revealed a significant increase in the Ce, Zr,
and Rh surface concentrations after washing with a weak oxalic acid
solution (0.1 M). This was ascribed to the uncovering of the TWC
washcoat surface on removal of the P crust.165–167 Figure 3.22(b)
shows XRD patterns of the aged TWC samples and TWC samples
regenerated with EDTA, citric, and oxalic acids. It can be observed
that the peaks assigned to (Mg,Ca,Zn)3(PO4)2, Ce and Al phos-
phates disappear after washing with oxalic acid. The BET area and
pore volume of the washed catalyst increased, which implies that the
removal of large quantities of P-containing compounds led to the
unclogging of fouled washcoat pores.165,167 Thus, diffusional resist-
ance caused by the blocking of pore mouths might be overcome and
the diffusion rate of gases at the newly exposed catalytic sites can be
significantly enhanced. The latter was confirmed by in situ DRIFTS
CO and NO chemisorption studies, which indicated significant
enhancement in the concentration of adsorbed CO and NO on Pd
and Rh surfaces after regeneration with oxalic acid.165
All regeneration procedures applied led to the partial recovery
of catalytic activity of the TWC (CO, CxHy, and NOx conversions)
under real exhaust-gas conditions (dynamometer tests), and
improvements in the oxygen storage capacity, due to the removal
of large amounts of contaminants from the catalyst surface.168
However, oxalic acid washing was found to be the most effective in
improving the oxygen storage and release rates evaluated by pulse
(Fig. 3.23(a)) and step-change OSC measurements, a result that is
strongly related to its efficacy in removing P-containing spe-
cies.164–167 The most significant improvements in the catalytic activ-
ity of an aged TWC for CO (Fig. 3.23(b)), CxHy, and NOx
conversions were also obtained after oxalic acid washing.166,167,169
Similar results were also obtained by Shim et al.172 by evaluating

b1469_Ch-03.indd 209 4/8/2013 12:29:11 PM


b1469 Catalysis by Ceria and Related Materials

210 A. M. Efstathiou and S. Y. Christou

200
Aged
Oxalic acid

OSC (µmol-O/gcat)
150 Citric acid
EDTA

100

50

0
450 550 650 750
Temperature ( o C)
(a)

100
Aged
Oxalic acid
80
Citric acid
EDTA
XCO (%)

60

40

20

0
200 250 300 350 400 450
Temperature ( o C)
(b)

Figure 3.23 (a) OSC and (b) XCO values for aged TWC samples (83,000 km) and
TWC samples regenerated after using oxalic acid, citric acid, and EDTA. Adapted
with permission from Christou.169 Copyright 2006 Central Library of the University
of Cyprus.

acetic, citric, and oxalic acid washing procedures for regaining the
oxidation activity of spent TWCs with 120,000 and 160,000 km mile-
age towards VOCs. The acid-treated TWCs were found to be more
active than the untreated ones, with the oxalic acid treatment being
the most suitable in improving the oxidation of aromatic
hydrocarbons.172

b1469_Ch-03.indd 210 4/8/2013 12:29:11 PM


b1469 Catalysis by Ceria and Related Materials

Investigation of the Oxygen Storage and Release Kinetics of TWCs 211

An ideal TWC regeneration treatment should be able to reverse


both thermal and chemical deactivation effects. Based on this
rationale, a treatment combining both thermal and liquid-chlorine
procedures was performed on an aged commercial TWC (30,000
km mileage), and its efficacy in improving metal dispersion and at
the same time removing various contaminants from the catalyst sur-
face was evaluated.169 The regeneration procedure consisted of an
initial submerging of the catalyst sample in a 0.5 M HCl solution for
30 min at 25 °C while stirring, drying at 100 °C overnight, and finally
exposing the catalyst to 500 °C in a 20 vol% O2/N2 gas flow for 2 h.
Hydrogen chemisorption experiments indicated a significant loss
(> 80%) of the available active surface of Pd and Rh metals in an
aged TWC compared to when fresh. Scanning electron microscopy/
energy dispersive X-ray (SEM/EDX) analyses showed a large num-
ber of Pd crystallites of size 50–100 nm. Moreover, the BET surface
area of the fresh TWC was found to drop significantly (∼60%) with
increasing mileage (0 to 30,000 km). These changes were attributed
to the collapse of pore structure due to fouling of the washcoat
pores by contaminants and sintering of the support via pore closure.
The regeneration treatment was found to increase metal dispersion
by 40–77%, with the number density of crystallites of up to 100 nm
in size reduced by ∼60% when compared to the aged TWC. The
BET area was found to increase only slightly, probably due to the
severe restructuring of the washcoat material. Nonetheless, the num-
ber of small pores (< 10 nm) was found to have increased in the
treated sample due to the complete removal of P (detected on
the outermost surface of the aged TWC) and the partial removal of
S (25–40%) deposited deeper in the washcoat. Catalytic activity tests
revealed a substantial improvement in catalyst performance after
regeneration. More precisely, T50 (the temperature at which 50%
conversion is attained) for CO and HC pollutants was found to have
decreased by ∼80 °C, while remarkable improvements in NOx
conversion were also achieved.
The oxalic acid regeneration treatment, which was proven to be
very effective in rejuvenating the catalytic activity of severely aged

b1469_Ch-03.indd 211 4/8/2013 12:29:11 PM


b1469 Catalysis by Ceria and Related Materials

212 A. M. Efstathiou and S. Y. Christou

Table 3.6 CO, NOx, and HC emission results (g·km−1) from NEDC tests performed
over aged and regenerated commercial TWCs. (Adapted from Christou et al.168)
P42 M156 A165
Emissions
(g·km−1) Aged Regenerated Aged Regenerated Aged Regenerated
CO 1.90 1.70 4.50 2.30 2.20 2.00
NOx 0.33 0.13 1.04 0.52 0.44 0.11
HC 0.28 0.23 0.47 0.32 0.38 0.26

three-way catalysts, was tested on full-scale commercial TWC converters


(42,500, 156,000, and 165,000 km mileage).168 The efficiency of the
regeneration procedure was found to depend on catalyst mileage and
the degree of contamination. Table 3.6 shows emission results (g·km−1)
obtained during New European Driving Cycle (NEDC) tests per-
formed over aged and regenerated commercial TWCs. A reduction of
up to about 50% in CO, 75% in NOx, and 32% in HC emissions were
attained under warm engine driving conditions, following regenera-
tion of the aged TWC converters. It was thus proven that expanding
the durability of deactivated commercial TWCs is feasible after apply-
ing washing treatments using weak oxalic acid solutions. This method-
ology appears at least to be a very promising starting point in recovering
to a large degree the catalytic performance of commercial TWCs.168

Acknowledgments
The European Regional Development Fund, the Republic of
Cyprus, the Cyprus Research Promotion Foundation, and the
Research Committee of the University of Cyprus are gratefully
acknowledged for their financial support through the project
DIDAKTOR/DISEK/0308/33.

References
1. Herz, R.K., Ind. Eng. Chem. Prod. Res. Dev., 20 (1981) 451–457.
2. Herz, R.K., Kiela, J.B., Sell, J.A., Ind. Eng. Chem. Prod. Res. Dev., 22 (1983)
387–396.

b1469_Ch-03.indd 212 4/8/2013 12:29:11 PM


b1469 Catalysis by Ceria and Related Materials

Investigation of the Oxygen Storage and Release Kinetics of TWCs 213

3. Herz, R.K., Sell, J.A., J. Catal., 94 (1985) 166–174.


4. Herz, R.K., Shinouskis, E.J., Ind. Eng. Chem. Prod. Res. Dev., 24 (1985)
385–390.
5. Herz, R.K., Stud. Surf. Sci. Catal., 30 (1987) 427–444.
6. Yao, H.C., Yu Yao, Y.F., J. Catal., 86 (1984) 254–265.
7. Su, E.C., Montreuil, C.N., Rothschild, W.G., Appl. Catal., 17 (1985)
75–86.
8. Su, E.C., Rothschild, W.G., J. Catal., 99 (1986) 506–510.
9. Su, E.C., Rothschild, W.G., Yao, H.C., J. Catal., 118 (1989) 111–124.
10. Engler, B., Koberstein, E., Schubert, P, Appl. Catal., 48 (1989) 71–92.
11. Lööf, P., Kacemo, B., Keck, K.-E., J. Catal., 118 (1989) 339–348.
12. Kacimi, S., Barbier, J., Jr., Taha, R., Duprez D., Catal. Lett., 22 (1993)
343–350.
13. Trovarelli, A., Catal. Rev., 38 (1996) 439–520.
14. Fornasiero, P., Balducci, G., Di Monte, R., Kašpar, J., Sergo, V.,
Gubitosa, G., Ferrero, A., Graziani, M., J. Catal., 164 (1996) 173–183.
15. Fornasiero, P., Balducci, G., Kašpar, J., Meriani, S., Di Monte, R.,
Graziani, M., Catal. Today, 29 (1996) 47–52.
16. Kašpar, J., Fornasiero, P., Graziani, M., Catal. Today, 50 (1999)
285–298.
17. Kašpar, J., Graziani, M., Fornasiero, P., in Handbook on the Physics and
Chemistry of Rare Earths, Vol. 29, eds. K.A. Gschneidner Jr., L. Eyring,
Elsevier Science B.V., Amsterdam, (2000), pp. 159–267.
18. Trovarelli, A., in Catalysis by Ceria and Related Materials, ed. A. Trovarelli,
Imperial College Press, London, (2002) pp. 15–50.
19. Duprez, D., Descorme, C., in Catalysis by Ceria and Related Materials, ed.
A. Trovarelli, Imperial College Press, London, (2002), pp. 243–280.
20. Kašpar, J., Fornasiero, P., J. Solid State Chem., 171 (2003) 19–29.
21. Sugiura, M., Catal. Surv. Asia, 7 (2003) 77–87.
22. Matsumoto, S., Catal. Today, 90 (2004) 183–190.
23. Möller, R., Votsmeier, M., Onder, C., Guzzella, L., Gieshoff, J., Appl.
Catal. B, Environ., 91 (2009) 30–38.
24. Martin, D., Duprez, D., J. Phys. Chem., 100 (1996) 9429–9438.
25. Descorme, C., Duprez, D., Appl. Catal. A, Gen., 202 (2000) 231–241.
26. Bedrane, S., Descorme, C., Duprez, D., Appl. Catal. A, Gen., 289
(2005) 90–96.

b1469_Ch-03.indd 213 4/8/2013 12:29:11 PM


b1469 Catalysis by Ceria and Related Materials

214 A. M. Efstathiou and S. Y. Christou

27. Bedrane, S., Descorme, C., Duprez, D., Catal. Today, 75 (2002)
401–405.
28. Di Monte, R., Kašpar, J., Top. Catal., 28 (2004) 47–57.
29. Boaro, M., Giordano, F., Recchia, S., Dal Santo, V., Giona, M.,
Trovarelli, A., Appl. Catal. B, Environ., 52 (2004) 225–237.
30. Aneggi, E., Boaro, M., De Leitenburg, C., Dolcetti, G., Trovarelli, A.,
J. Alloys Compd., 408–412 (2006) 1096–1102.
31. Koltsakis, G.C., Stamatelos, A.M., Prog. Energy Combust. Sci., 23 (1997)
1–39.
32. Costa, C.N., Christou, S.Y., Georgiou, G., Efstathiou, A.M., J. Catal.,
219 (2003) 259–272.
33. Yamada, K., Tanaka, H., Yamamoto, M., SAE Technical Paper 970464
(1997).
34. Lambrou, P.S., Costa, C.N., Christou, S.Y., Efstathiou, A.M., Appl.
Catal. B, Environ., 54 (2004) 237–250.
35. Hickey, N., Fornaciero, P., Kašpar, J., Gatica, J.M., Bernal, S., J. Catal.,
200 (2001) 181–193.
36. Duprez, D., Descorme, C., Birchem, T., Rohart, E., Top. Catal., 16/17
(2004) 49–56.
37. Christou, S.Y., Costa, C.N., Efstathiou, A.M., Top. Catal., 30/31 (2004)
325–331.
38. Salomons, S., Hayes, R.E., Votsmeier, M., Drochner, A., Vogel, H.,
Malmberg, S., Gieshoff, J., Appl. Catal. B, Environ., 70 (2007) 305–313.
39. Li, C., Sakata, Y., Arai, T., Domen, K., Maruya, K.-I., Onishi, T., J. Chem.
Soc., Faraday T. 1, 85 (1989) 929–943.
40. Li, C., Sakata, Y., Arai, T., Domen, K., Maruya, K.-I., Onishi, T., J. Chem.
Soc., Faraday T. 1, 85 (1989) 1451–1461.
41. Holmgren, A., Andersson, B., Duprez, D., Appl. Catal. B, Environ., 22
(1999) 215–230.
42. Song, Z., Liu, W., Nishigushi, H., Catal. Commun., 8 (2007) 725–730.
43. Bernal, S., Calvino, J.J., Gatica, J.M., Cartes, C.L., Pintado, J.M. in
Catalysis by Ceria and Related Materials, ed. A. Trovarelli, Imperial
College Press, London, (2002), pp. 85–168.
44. Boaro, M., de Leitenburg, C., Dolcetti, G., Trovarelli, A., J. Catal., 193
(2000) 338–347.

b1469_Ch-03.indd 214 4/8/2013 12:29:11 PM


b1469 Catalysis by Ceria and Related Materials

Investigation of the Oxygen Storage and Release Kinetics of TWCs 215

45. Descorme, C., Taha, R., Mouaddib-Moral, N., Duprez, D., Appl. Catal.
A, Gen., 223 (2002) 287–299.
46. Aneggi, E., Boaro, M., de Leitenburg, C., Dolcetti, G., Trovarelli, A.,
Catal. Today, 112 (2006) 94–98.
47. Meiqing, S., Xinquan, W., Yuan, W., Duan, W., Minwei, Z., Jun, W.,
J. Rare Earths, 25 (2007) 48–52.
48. Zhao, M., Shen, M., Wang, J., J. Catal., 248 (2007) 258–267.
49. Wang, B., Weng, D., Wu, X., Fan, J., Catal. Today, 153 (2010) 111–117.
50. Wang, Q., Li, G., Zhao, B., Zhou, R., J. Mol. Catal. A, Chem., 339 (2011)
52–60.
51. Wang, Q., Li, G., Zhao, B., Zhou, R., Fuel, 90 (2011) 3047–3055.
52. Li, G., Wang, Q., Zhao, B., Zhou, R., Fuel, 92 (2012) 360–368.
53. Fornasiero, P., Fonda, E., Di Monte, R., Vlaic, G., Kašpar, J., Graziani,
M., J. Catal., 187 (1999) 177–185.
54. Vidal, H., Bernal, S., Kašpar, J., Pijolat, M., Perrichon, V., Blanco, G.,
Pintado, J.M., Baker, R.T., Colon, G., Fally, F., Catal. Today, 54 (1999)
93–100.
55. Nibbelke, R.H., Nievergerd, A.J.L., Hoebnik, J.H.B.J., Marin, G.B.,
Appl. Catal. B, Environ., 19 (1998) 245–259.
56. Sakamoto, Y., Kizaki, K., Motohiro, T., Yokota, Y., Sobukawa, H.,
Uenishi, M., Tanaka, H., Sugiura, M., J. Catal., 211 (2002) 157–164.
57. Miyamoto, K., Takebayashi, H., Ishihara, T., Kido, H., Hatamura, K.,
SAE Technical Paper 2002–01–1094 (2002).
58. Holmgren, A., Andersson, B., J. Catal., 178 (1998) 14–25.
59. Nievergeld, A.J.L., PhD thesis, Eindhoven University of Technology
(1998).
60. Tsinoglou, D.N., Koltsakis, G.C., Peyton Jones, J.C., Ind. Eng. Chem.
Res., 41 (2002) 1152–1165.
61. Larese, C., Galisteo, F.C., Granados, M.L., Mariscal, R., Fierro, J.L.G.,
Lambrou, P.S., Efstathiou, A.M., J. Catal., 226 (2004) 443–456.
62. Martin, D., Kaur, P., Duprez, D., Gaigneaux, E., Ruiz, P., Delmon, B.,
Catal. Today, 32 (1996) 329–336.
63. Ezin, A.N., Tsilidilkovski, V.I., Kurumchin, E.K., Solid State Ionics, 84
(1996) 105–112.
64. Doornkamp, C., Clement, M., Ponec, V., J. Catal., 182 (1999) 390–399.

b1469_Ch-03.indd 215 4/8/2013 12:29:11 PM


b1469 Catalysis by Ceria and Related Materials

216 A. M. Efstathiou and S. Y. Christou

65. Madier, Y., Descorme, C., Le Govic, A.M., Duprez, D., J. Phys. Chem. B,
103 (1999) 10999–11006.
66. Descorme, C., Madier, Y., Duprez, D., J. Catal., 196 (2000) 167–173.
67. Den Otter, M.W., Boukamp, B.A., Bouwmeester, H.J.M., Solid State
Ionics, 139 (2001) 89–94.
68. Larese, C., Galisteo, F.C., Granados, M.L., Lopez, R.M., Fierro, J.L.G.,
Lambrou, P.S., Efstathiou, A.M., Appl. Catal. B, Environ., 48 (2004)
113–123.
69. Larese, C., Granados, M.L., Mariscal, R., Fierro, J.L.G., Lambrou, P.,
Efstathiou, A.M., Appl. Catal. B, Environ., 59 (2005) 13–25.
70. Granados, M.L., Galisteo, F.C., Lambrou, P.S., Mariscal, R., Sanz, J.,
Sobrados, I., Fierro, J.L.G., Efstathiou, A.M., J. Catal., 239 (2006)
410–421.
71. Lambrou, P.S., Efstathiou, A.M., J. Catal., 240 (2006) 182–193.
72. Christou, S.Y., Bradshaw, H., Butler, C., Darab, J., Efstathiou, A.M.,
Top. Catal., 52 (2009) 2013–2018.
73. Christou, S.Y., Alvarez, M.C.G., Fierro, J.L.G., Efstathiou, A.M., Appl.
Catal. B, Environ., 106 (2011) 103–113.
74. Cunningham, J., Farrell, F., Gibson, C., McCarthy, J., Catal. Today, 50
(1999) 429–443.
75. Holmgren, A., Duprez, D., Andersson, B., J. Catal., 182 (1999) 441–448.
76. Ciuparu, D., Bozon-Verduraz, F., Pfefferle, L., J. Phys. Chem. B, 106
(2002) 3434–3442.
77. Jerdev, D.I., Kim, J., Batzill, M., Koel, B.E., Surf. Sci. Lett., 498 (2002)
L91–L96.
78. Galdikas, A., Duprez, D., Descorme, C., Appl. Surf. Sci., 236 (2004)
342–355.
79. Galdikas, A., Descorme, C., Duprez, D., Solid State Ionics, 166 (2004)
147–155.
80. Dong, F., Suda, A., Tanade, T., Nagai, Y., Sobukawa, H., Shinjoh, H.,
Sugiura, M., Descorme, C., Duprez, D., Catal. Today, 90 (2004)
223–229.
81. Lin, R., Inmaculada, R.-R., Belen, B.-B., Antonio, G.-R., Sun, G., Xin, Q.,
Chinese J. Catal., 27 (2006) 109–114.
82. Galdikas, A., Descorme, C., Duprez, D., Mater. Sci., 13 (2007)
193–205.

b1469_Ch-03.indd 216 4/8/2013 12:29:11 PM


b1469 Catalysis by Ceria and Related Materials

Investigation of the Oxygen Storage and Release Kinetics of TWCs 217

83. Vagia, E.Ch., Lemonidou, A.A., J. Catal., 269 (2010) 388–396.


84. Boreskov, G.K., Adv. Catal., 15 (1965) 285–339.
85. Winter, E.R.S., J. Chem. Soc., 1 (1968) 2889–2902.
86. Novakova, J., Catal. Rev., 4 (1970) 77–113.
87. Pushkarev, V.V., Kovalchuk, V.I., d’Itri, J.L., J. Phys. Chem. B, 108
(2004) 5341–5348.
88. Li, C., Domen, K., Maruya, K.-I., Onishi, T., J. Am. Chem. Soc., 111
(1989) 7683–7687.
89. Li, C., Domen, K., Maruya, K.-I., Onishi, T., J. Catal., 123 (1990)
436–442.
90. Rossignol, S., Gerard, F., Duprez, D., J. Mater. Chem., 9 (1999)
1615–1620.
91. Wu, Z., Li, M., Howe, J., Meyer, H.M., III, Overbury, S.H., Langmuir,
26 (2010) 16595–16606.
92. Kalamaras, C.M., Americanou, S., Efstathiou, A.M., J. Catal., 279
(2011) 287–300.
93. Efstathiou, A.M., Verykios, X.E., Appl. Catal. A, Gen., 151 (1997)
109–166.
94. Kramer, R., Andre, M., J. Catal., 58 (1979) 287–295.
95. Peil, K.P., Goodwin, J.G., Jr., Marcelin, G., J. Catal., 131 (1991) 143–155.
96. Sadovskaya, E.M., Ivanova, Y.A., Pinaeva, L.G., Grasso, G., Kuznetsova,
T.G., van Veen, A., Sadykov, V.A., Mirodatos C., J. Phys. Chem. A, 111
(2007) 4498–4505.
97. Klier, K., Kučera, E., J. Phys. Chem. Solids, 27 (1996) 1087–1095.
98. Happel, J., Walter, E., Lecourtier, Y., J. Catal., 123 (1990) 12–20.
99. Williamson, W.B., Perry, J., Gandhi, H.S., Bomback, J.L., Appl. Catal.,
15 (1985) 277–292.
100. Tabata, T., Baba, K., Kawashima, H., Appl. Catal. B, Environ., 7 (1995)
19–32.
101. Angelidis, T.N., Sklavounos, S.A., Appl. Catal. A, Gen., 133 (1995)
121–132.
102. Angelidis, T.N., Koutlemani, M.M., Sklavounos, S.A., Lioutas, Ch.B.,
Voulgaropoulos, A., Papadakis, V.G., Emons, H., Stud. Surf. Sci.
Catal., 116 (1998) 155–164.
103. Battistoni, C., Cantelli, V., Debenedetti, M., Kačiulis, S., Mattogno, G,
Napoli, A., Appl. Surf. Sci., 144–145 (1999) 390–394.

b1469_Ch-03.indd 217 4/8/2013 12:29:11 PM


b1469 Catalysis by Ceria and Related Materials

218 A. M. Efstathiou and S. Y. Christou

104. Bartholomew, C.H., Appl. Catal. A, Gen., 212 (2001) 17–60.


105. Rokosz, M.J., Chen, A.E., Lowe-Ma, C.K., Kucherov, A.V., Benson, D.,
Paputa Peck, M.C., McCabe, R.W., Appl. Catal. B, Environ., 33 (2001)
205–215.
106. Uy, D., O’Neill, A.E., Xu, L., Weber, W.H., McCabe, R.W., Appl. Catal.
B, Environ., 41 (2003) 269–278.
107. Fernandez-Ruiz, R., Furio, M., Cabello Galisteo, F., Larese, C., Lopez
Granados, M., Fierro, J.L.G., Anal. Chem., 74 (2002) 5463–5469.
108. Xu, L., Guo, G., Uy, D., O’Neil, A.E., Weber, W.H., Rokosz, M.J.,
McCabe, R.W., Appl. Catal. B, Environ., 50 (2004) 113–125.
109. Kroger, V., Hietikko, M., Lassi, U., Ahola, J., Kallinen, K., Laitinen, R.,
Keiski, R.L., Top. Catal., 30/31 (2004) 469–473.
110. Kroger, V., Lassi, U., Kynkaanniemi, K., Suopanki, A., Keiski, R.L.,
Chem. Eng. J., 120 (2006) 113–118.
111. Larese, C., Lopez Granados, M., Cabello Galisteo, F., Mariscal, R.,
Fierro, J.L.G., Furio, M., Fernandez Ruiz, R., Appl. Catal. B, Environ.,
62 (2005) 132–143.
112. Larese, C., Cabello Galisteo, F., Lopez Granados, M., Mariscal, R.,
Fierro, J.L.G., Furio, M., Fernandez Ruiz, R., Appl. Catal. B, Environ.,
40 (2003) 305–317.
113. Lopez Granados, M., Larese, C., Cabello Galisteo, F., Mariscal, R.,
Fierro, J.L.G., Fernandez-Ruiz, R., Sanguino, R., Luna, M., Catal.
Today, 107–108 (2005) 77–85.
114. Shelef, M., Graham, G.W., McCabe, R.W., in Catalysis by Ceria and
Related Materials, ed. A. Trovarelli, Imperial College Press, London,
(2002), pp. 343–376.
115. Lopez Granados, M., Cabello Galisteo, F., Lambrou, P.S., Alifanti, M.,
Mariscal, R., Gurbani, A., Sanz, J., Sobrados, I., Efstathiou, A.M.,
Fierro, J.L.G., Top. Catal., 42/43 (2007) 443–447.
116. Granados, M.L., Gurbani, A., Mariscal, R., Fierro, J.L.G., J. Catal., 256
(2008) 172–185.
117. Heo, I., Choung, J.W., Kim, P.S., Nam, I.-S., Song, Y.I., In, C.B.,
Yeo, G.K., Appl. Catal. B, Environ., 92 (2009) 114–125.
118. Christou, S.Y., Garcia-Rodriguez, S., Fierro, J.L.G., Efstathiou, A.M.,
Appl. Catal. B, Environ., 111–112 (2012) 233–245.
119. Christou, S.Y., Efstathiou, A.M., Top. Catal., 42/43 (2007) 351–355.

b1469_Ch-03.indd 218 4/8/2013 12:29:11 PM


b1469 Catalysis by Ceria and Related Materials

Investigation of the Oxygen Storage and Release Kinetics of TWCs 219

120. Lambrou, P.S., Savva, P.G., Fierro, J.L.G., Efstathiou, A.M., Appl. Catal.
B, Environ., 76 (2007) 375–385.
121. Fernandes, D.M., Scofield, C.F., Neto, A.A., Baldini Cardoso, M.J.,
Zanon Zotin, F.M., Process Saf. Environ., 87 (2009) 315–322.
122. Fernandes, D.M., Scofield, C.F., Neto, A.A., Baldini Cardoso, M.J.,
Zanon Zotin, F.M., Chem. Eng. J., 160 (2010) 85–92.
123. Neyestanaki, A.K., Klingstedt, F., Salmi, T., Murzin, D.Yu., Fuel, 83
(2004) 395–408.
124. Hori, C.E., Permana, H., Simon Ng, K.Y., Brenner, A., More, K.,
Rahmoeller, K.M., Belton, D., Appl. Catal. B, Environ., 16 (1998)
105–117.
125. Graham, G.W., Jen, H.-W., Chun, W., McCabe, R.W., J. Catal., 182
(1999) 228–233.
126. Inglesias-Juez, A., Martínez-Arias, A., Fernández-García, M., J. Catal.,
221 (2004) 148–161.
127. Lassi, U., Polvinen, R., Suhonen, S., Kallinen, K., Savimaki, A., Harkonen,
M., Valden, M., Keiski, R.L., Appl. Catal. A, Gen., 263 (2004) 241–248.
128. Jen, H.-W., Graham, G.W., Chun, W., McCabe, R.W., Cuif, J.-P.,
Deutsch, S.E., Touret, O., Catal. Today, 50 (1999) 309–328.
129. Kozlov, A.I., Kim, D.H., Yezerets, A., Andersen, P., Kung, H.H.,
Kung, M.C., J. Catal., 209 (2002) 417–426.
130. Liotta, L.F., Longo, A., Pantaleo, G., Di Carlo, G., Martorana, A.,
Cimino, S., Russo, G., Deganello, G, Appl. Catal. B, Environ., 90 (2009)
470–477.
131. Yashima, M., Arashi, H., Kakihana, M., Yoshimura, M., J. Am. Ceram.
Soc., 77 (1994) 1067–1071.
132. Kenevey, K., Valdivieso, F., Soustelle, M., Pijolat, M., Appl. Catal. B,
Environ., 29 (2001) 93–101.
133. Bedrane, S., Descorme, C., Duprez, D., Catal. Today, 73 (2002) 233–238.
134. Hori, C.E., Brenner, A., Simon Ng, K.Y., Rahmoeller, K.M., Belton, D.,
Catal. Today, 50 (1999) 299–308.
135. Fan, J., Wu, X., Wu, X., Liang, Q., Ran, R., Weng, D., Appl. Catal. B,
Environ., 81 (2008) 38–48.
136. Bernal, S., Blanco, G., Calvino, J.J., Gatica, J.M., Pérez-Omil, J.A.,
Pintado, J.M., Top. Catal., 28 (2004) 31–45.
137. Fan, J., Wu, X., Ran, R., Weng, D., Appl. Surf. Sci., 245 (2005) 162–171.

b1469_Ch-03.indd 219 4/8/2013 12:29:11 PM


b1469 Catalysis by Ceria and Related Materials

220 A. M. Efstathiou and S. Y. Christou

138. Liotta, L.F., Longo, A., Macaluso, A., Martorana, A., Pantaleo, G.,
Venezia, A.M., Deganello, G, Appl. Catal. B, Environ., 48 (2004) 133–149.
139. Kašpar, J., Di Monte, R., Fornasiero, P., Graziani, M., Bradshaw, H.,
Norman, C., Top. Catal., 16/17 (2001) 83–87.
140. Hickey, N., Fornasiero, P., Di Monte, R., Kašpar, J., Graziani, M.,
Dolcetti, G., Catal. Lett., 72 (2001) 45–50.
141. Fally, F., Perrichon, V., Vidal, H., Kašpar, J., Blanco, G., Pintado, J.M.,
Bernal, S., Colon, G., Daturi, M., Lavalley, J.C., Catal. Today, 59 (2000)
373–386.
142. Vidal, H., Kašpar, J., Pijolat, M., Colon, G., Bernal, S., Gordón, A.,
Perrichon, V., Fally, F., Appl. Catal. B, Environ., 27 (2000) 49–63.
143. Vidal, H., Kašpar, J., Pijolat, M., Colon, G., Bernal, S., Gordón, A.,
Perrichon, V., Fally, F., Appl. Catal. B, Environ., 30 (2001) 75–85.
144. Otsuka-Yao, S., Morikawa, H., Izu, N., Okuda, K., J. Japan Inst. Metals,
59 (2003) 1237–1246.
145. Otsuka-Yao-Matsuo, S., Omata, T., Izu, N., Kishimoto, H., J. Solid State
Chem., 138 (1998) 47–54.
146. Omata, T., Kishimoto, H., Otsuka-Yao-Matsuo, S., Ohtori, N.,
Umesaki, N., J. Solid State Chem., 147 (1999) 573–583.
147. Kishimoto, H., Omata, T., Otsuka-Yao-Matsuo, S., Ueda, K., Hosono,
H., Kawazoe, H., J. Alloys Compd., 312 (2000) 94–103.
148. Di Monte, R., Kašpar, J., J. Mater. Chem., 15 (2005) 633–648.
149. Nagai, Y., Yamamoto, T., Tanaka, T., Yoshida, S., Nonaka, T.,
Okamoto, T., Suda, A., Sugiura, M., Top. Catal., 47 (2008) 137–147.
150. Yeste, M.P., Hernández, J.C., Bernal, S., Blanco, G., Calvino, J.J., Pérez-
Omil, J.A., Pintado, J.M., Chem. Mater., 18 (2006) 2750–2757.
151. Bernal, S., Blanco, G., Calvino, J.J., Hernández, J.C., Pérez-Omil, J.A.,
Pintado, J.M., Yeste, M.P., J. Alloys Compd., 451 (2008) 521–524.
152. Yeste, M.P., Hernández, J.C., Trasobares, S., Bernal, S., Blanco, G.,
Calvino, J.J., Pérez-Omil, J.A., Pintado, J.M., Chem. Mater., 20 (2008)
5107–5113.
153. Yeste, M.P., Hernández, J.C., Bernal, S., Blanco, G., Calvino, J.J., Pérez-
Omil, J.A., Pintado, J.M., Catal. Today, 141 (2009) 409–414.
154. Hernández, J.C., Hungría, A.B., Pérez-Omil, J.A., Trasobares, S.,
Bernal, S., Midgley, P.A., Alavi, A., Calvino, J.J., J. Phys. Chem. C, 111
(2007) 9001–9004.

b1469_Ch-03.indd 220 4/8/2013 12:29:11 PM


b1469 Catalysis by Ceria and Related Materials

Investigation of the Oxygen Storage and Release Kinetics of TWCs 221

155. Lemaux, S., Bensaddik, A., van der Eerden, A.M.J., Bitter, J.H.,
Koningsberger, D.C., J. Phys. Chem. B, 105 (2001) 4810–4815.
156. Nagai, Y., Yamamoto, T., Tanaka, T., Yoshida, S., Nonaka, T., Okamoto,
T., Suda, A., Sugiura, M., Catal. Today, 74 (2002) 225–234.
157. Birgersson, H., Boutonnet, M., Järås, S., Eriksson, L., Top. Catal.,
30/31 (2004) 433–437.
158. Birgersson, H., Eriksson, L., Boutonnet, M., Järås, S., Appl. Catal. B,
Environ., 54 (2004) 193–200, and references therein.
159. Birgersson, H., Boutonnet, M., Klingstedt, F., Murzin, D.Y., Stefanov, P.,
Naydenov, A., Appl. Catal. B, Environ., 65 (2006) 93–100.
160. Daley, R.A., Christou, S.Y., Efstathiou, A.M., Anderson, J.A., Appl.
Catal. B, Environ., 60 (2005) 117–127.
161. Anderson, J.A., Daley, R.A., Christou, S.Y., Efstathiou, A.M., Appl.
Catal. B, Environ., 64 (2006) 189–200.
162. Angelidis T.N., Papadakis, V.G., Appl. Catal. B, Environ., 12 (1997)
193–206.
163. Darr, S.T., Choksi, R.A., Hubbard, C.P., Johnson, M.D., McCabe, R.W.,
SAE Technical Paper 2000–01–1881 (2000).
164. Lambrou, P.S., Christou, S.Y., Fotopoulos, A.P., Foti, F.K., Angelidis,
T.N., Efstathiou, A.M., Appl. Catal. B, Environ., 59 (2005) 1–11.
165. Christou, S.Y., Birgersson, H., Fierro J.L.G., Efstathiou, A.M., Environ.
Sci. Technol., 40 (2006) 2030–2036.
166. Christou, S.Y., Efstathiou, A.M., Top. Catal., 42/43 (2007) 415–419.
167. Christou, S.Y., Birgersson, H., Efstathiou, A.M., Appl. Catal. B, Environ.,
71 (2007) 185–198.
168. Christou, S.Y., Gåsste, J., Karlsson, H.L., Fierro, J.L.G., Efstathiou,
A.M., Top. Catal., 52 (2009) 2029–2034.
169. Christou, S.Y., PhD Thesis, University of Cyprus (2006).
170. Subramanian, B., Christou, S.Y., Efstathiou, A.M., Dionysiou, D.,
J. Hazard. Mater., 186 (2011) 999–1006.
171. Kim, S.C., Nahm, S.W., Shim, W.G., Seo, S.G., Lee, J.W., J. Nanosci.
Nanotechnol., 8 (2008) 5398–5403.
172. Shim, W.G., Jung, S.C., Seo, S.G., Kim, S.C., Catal. Today, 164 (2011)
500–506.
173. Lambrou, P.S., Polychronopoulou, K., Petallidou, K.C., Efstathiou, A.M.,
Appl. Catal. B, Environ., 111–112 (2012) 349–359.

b1469_Ch-03.indd 221 4/8/2013 12:29:11 PM


b1469 Catalysis by Ceria and Related Materials

This page intentionally left blank

b1469_Ch-03.indd 222 4/8/2013 12:29:11 PM


b1469 Catalysis by Ceria and Related Materials

CHAPTER 4

INTERACTION OF NITROGEN OXIDES


WITH CERIA-BASED MATERIALS
Avelina García-García and Agustin Bueno-López
Department of Inorganic Chemistry,
University of Alicante, Ap. 99 E-03080, Alicante, Spain

4.1 Introduction
Nitrogen is able to form no fewer than eight oxides1 and three of
these gases (N2O, NO, and NO2) are of special interest due to their
environmental and health impact.
Nitrous oxide (N2O) has long been considered a relatively
harmless gas and its atmospheric emissions have not received
much attention from scientists, engineers, or politicians. However,
there is growing concern since N2O is a harmful gas in our environ-
ment contributing to the greenhouse effect and the depletion of
the ozone layer. Although N2O is not the major contributor to
global warming it is much more potent than the other two most
common anthropogenic greenhouse gases, CO2 and CH4. Due to
its long lifespan of approximately 150 years in the atmosphere,
N2O has 310 and 21 times the global-warming potential (GWP) of
CO2 and CH4, respectively.2,3 The estimated contribution of anthro-
pogenic N2O emissions to the atmosphere is 4.7–7 million tons per
year (1 ton = 907.2 kg) and some scientists point out the need for
a 70–80% reduction in human emissions to achieve a null balance
between N2O emissions and N2O abatement provided naturally by
the planet.2,3 From the point of view of the greenhouse effect, the
required reduction in N2O emissions from anthropogenic sources

223

b1469_Ch-04.indd 223 4/8/2013 12:29:58 PM


b1469 Catalysis by Ceria and Related Materials

224 A. García-García and A. Bueno-López

is equivalent to the CO2 abatement committed to in the Kyoto


protocol.
Anthropogenic sources of N2O include adipic and nitric acid
production, fossil fuel and biomass combustion, land cultivation,
and vehicle emissions.2–6 N2O emissions from gasoline-powered
engines have been related to the aging of three-way catalysts (TWCs)
and N2O is also emitted as a by-product of Pt-based prototypic cata-
lysts for the selective catalytic reduction (SCR) of NOx with hydro-
carbons in diesel engine exhausts.7
N2O emissions can be controlled by catalytic decomposition,
since N2O is thermodynamically unstable and suitable catalysts
increase the decomposition rate to N2 and O2. A number of catalysts
based on bulk and supported mixed metal oxides and metal-loaded
zeolites have been implemented for N2O decomposition in adipic
acid plants (CuAl2O4/Al2O3 and Ag/Al2O3 by BASF, CoO–NiO/
ZrO2 by DuPont, and CuO/Al2O3 by Asahi Chemical)8 and in nitric
acid plants (CuO/Al2O3 by BASF, La0.8Ce0.2CoO3 by Johnson Matthey,
Co2AlO4/CeO2 by Yara International, and FeZSM-5 by Uhde).3
Other catalytic systems have been evaluated under simulated condi-
tions of adipic acid and nitric acid plants, such as yttrium-doped
zirconia;9 AB1-xB′xO3 perovskites with A = La or Ca, B = Mn or Fe,
and B′ = Cu or Ni;10 ABAl11O19 hexaaluminates with A = Ba or La and
B = Fe, Mn, Co, or Ni;11 γ-alumina-supported iron oxide;12 mayen-
ite;13 and noble metals/ceria.14–19 Metal catalysts, including Pt, Pd,
Ag, Au, Fe, and Ge, are generally active above 375°C and, of the
pure oxides, the highest activities are exhibited by the transition
metal oxides and by some rare earth oxides. A comparison of spe-
cific activities (per unit surface area) of various pure oxides revealed
that Rh2O3 has one of the best activities,2 and ceria promoters have
been demonstrated to enhance Rh2O3 activity significantly. Catalysts
of rhodium supported on Zr-, La-, or Pr-doped CeO2, zirconia–alu-
mina and Rh/CeO2–TiO2/monolithic catalysts have been reported
to be highly effective for low-temperature N2O decomposition.14–18
Nitric oxide (NO) and nitrogen dioxide (NO2) are emitted by
transportation and industrial plants. For example, they are pro-
duced during fossil fuel combustion in electrical power plants or

b1469_Ch-04.indd 224 4/8/2013 12:29:58 PM


b1469 Catalysis by Ceria and Related Materials

Interaction of Nitrogen Oxides with Ceria-Based Materials 225

gasoline and diesel combustion in vehicles. NO accounts for 95% of


all NO + NO2 emissions, but in the presence of air NO is progres-
sively oxidized to NO2, which is a stable oxide at low temperatures.
Due to the equilibrium:

NO + 1/2O2  NO2

the emissions of NO and NO2 are generically referred to as NOx.


NOx contributes to a variety of environmental and health prob-
lems such as acid rain, ground-level ozone formation, and lung dis-
ease. Epidemiological studies have revealed that the concentration
of NOx having a hazardous effect on someone in good health is
above 0.05 ppm for an exposure over 24 hours, and this value is
often exceeded in towns with dense traffic during peak hours.20
Catalysts are also widely used for NOx abatement both in station-
ary and mobile sources, and ceria-based oxides play a key role in
some of these catalysts, mainly in those used in vehicles. Conventional
TWCs used in gasoline vehicles for simultaneous CO, hydrocarbon,
and NOx removal combine noble metals with cerium-based oxides
for oxygen storage.21 Recently, cerium oxides have also been demon-
strated to be promising candidates for the regeneration of soot fil-
ters fitted in diesel vehicles,22–26 and NOx is involved in ceria-catalyzed
soot oxidation in a typical diesel exhaust environment. It has been
proposed that ceria has the potential to accelerate the oxidation
rate of soot due to its active oxygen and that the formation of such
active oxygen is initiated by NO2 in the gas phase.22 In some situa-
tions, a relation between the catalytic oxidation of NO to NO2 (NO2
being much more oxidizing than O2 and NO) and soot combus-
tion27 has been found.
In this chapter, two catalyzed reactions with practical relevance
have been selected to provide an overview of the interaction of
nitrogen oxides (N2O, NO, and NO2) with ceria-based materials.
The first is the catalytic decomposition of N2O on Rh/ceria catalysts,
which is relevant for N2O pollution control, and the other is the
ceria-catalyzed oxidation of NO to NO2, which is involved in soot
combustion in diesel exhausts.

b1469_Ch-04.indd 225 4/8/2013 12:29:58 PM


b1469 Catalysis by Ceria and Related Materials

226 A. García-García and A. Bueno-López

4.2 N2O Interaction with Ceria-Based Materials:


N2O Decomposition on Rh/Ceria Catalysts
Although N2O is thermodynamically unstable, this molecule is quite
stable at room temperature. The activation energy for thermal fis-
sion of the N-O bond is ca. 250–270 kJ/mol and temperatures above
625ºC are required to achieve measurable conversion2 according to:

2N2O → 2N2 + O2 AHο (25°C) = −163 kJ/mol

As mentioned in the introduction of this chapter, a number of


catalysts are able to accelerate N2O decomposition including ceria-
containing materials. Figure 4.1 shows N2O decomposition profiles
for different ceria samples and a reference Rh/γ-Al2O3 catalyst.
One of the main conclusions of these experiments is that the
Rh/ceria system is much more active for N2O decomposition than a
similar Rh/alumina catalyst (for details of the catalysts denoted by
0.5 wt% Rh/CeO2 and 0.5 wt% Rh/γ-Al2O3 in Fig. 4.1, refer to

100
N2O decomposed to N2 and O2 (%)

0.5 wt% Rh/Ce0.9Pr0.1O2 0.5 wt% Rh/γ-Al2O3

80
0.5 wt% Rh/CeO2

60

40
Ce0.9La0.1O2

20

CeO2 (calcined at 1000ºC)


0
100 200 300 400 500
Temperature (ºC)

Figure 4.1 Catalytic N2O decomposition of different materials at atmospheric


pressure. All data were obtained in steady-state conditions, in a fixed-bed reactor
coupled to a gas chromatograph. Gas mixture: 1000 ppm N2O/He; 12000 h−1. All
samples were powders and all cerias were calcined at 600ºC unless otherwise
indicated; Rh was loaded by wetness impregnation of a water solution of nitrate
followed by 500ºC calcination.

b1469_Ch-04.indd 226 4/8/2013 12:29:58 PM


b1469 Catalysis by Ceria and Related Materials

Interaction of Nitrogen Oxides with Ceria-Based Materials 227

N2O N2 + O2

N2O N2
O O

Rh Rh Rh

CeO2 CeO2-x CeO2-x


O Oxygen
2O vacancies

Rh Rh Rh

O2 N-N
N2O O

1/2O2 N2

Figure 4.2 N2O decomposition on Rh/ceria catalysts.

Parres-Esclapez et al.28). From electron spin resonance spectroscopy


(ESR) analysis of Rh/ceria catalysts, it was proposed that the low-
temperature N2O decomposition of these catalysts relies on electron
excess sites at microinterfaces between the dispersed rhodium com-
ponent and the ceria support.29 Experiments performed with iso-
topic 15N218O,28 among other techniques, evidenced that the ceria
supports, in the presence of rhodium, are actively involved in low-
temperature N2O decomposition while alumina is not. The reaction
mechanism proposed to explain N2O decomposition on Rh/ceria,
schematically represented in Fig. 4.2, is summarized in the following
sequence of steps:

Rh*−16O + 15N218O → Rh* + 15N2 + 18O16O Step 1


Rh* + 15N218O → Rh*−18O + 15N2 Step 2
Rh*−18O + 15N218O → Rh* + 15N2 + 18O2 Step 1 (repeated)
2Rh*−16O → 2Rh* + 16O2 Step 3
2Ce–OH + N2O → 2Ce* + N2 + H2O + O2 Step 4
Rh* + Ce*–16O → Rh*–16O + Ce* Step 5
Ce* + 15N218O → Ce*−18O + 15N2 Step 6
2Ce*−16O → 2Ce* + 16O2 Step 7
Rh*−16O + Ce*−16O → Rh* + Ce* + 16O2 Step 8

b1469_Ch-04.indd 227 4/8/2013 12:29:58 PM


b1469 Catalysis by Ceria and Related Materials

228 A. García-García and A. Bueno-López

where Rh*–O and Rh* represent oxidized and reduced sites of


rhodium, respectively, and Ce*–O and Ce* the oxidized and
reduced sites of ceria.
When N2O is in contact with a Rh/ceria or Rh/alumina catalyst,
if the temperature is high enough for N2O decomposition (above
200–250°C in the absence of inhibitors like H2O or NOx), N2O
reduces oxidized rhodium sites (Step 1). The re-oxidation of this
rhodium site can be carried out by another N2O molecule (Step 2),
and this sequence of steps (1 and 2) is the main reaction pathway
for Rh/alumina catalysts (called the Eley–Rideal mechanism). In
contrast, in Rh/ceria catalysts, the re-oxidation of rhodium sites can
be also accomplished by ceria oxygen (Step 5). The transfer of oxy-
gen from ceria to rhodium creates vacant sites on the ceria, and the
oxygen balance must be restored by re-oxidizing such vacant sites
with gas phase N2O (Step 6). N2O chemisorption on a ceria surface
not only occurs on vacant sites but also on hydroxyl groups, as in
Step 4 (this has been observed by diffuse reflectance IR Fourier-
transform spectroscopy (DRIFTS)30). The formation of high
amounts of 16−16O2 on 15N218O decomposition on Rh/ceria evidenced
that the recombination of two catalyst oxygen atoms to yield O2 also
occurs, as in Steps 3, 7, or 8.
In conclusion, several N2O decomposition pathways seem to
occur concomitantly on Rh/ceria catalysts, and all processes in Steps
1 to 8 participate in these N2O decomposition pathways. These pro-
cesses occurring on Rh/ceria are more effective than those based
exclusively on Steps 1 and 2 (occurring on Rh/alumina) because
active sites for N2O chemisorption and decomposition are not only
located on rhodium but also on ceria.
The better performance of the catalyst 0.5 wt% Rh/Ce0.9Pr0.1O2
(see Fig. 4.1) compared to the counterpart catalyst with bare ceria
(0.5 wt% Rh/CeO2) can be explained using this mechanism. It is
well known that the incorporation of foreign cations (of Pr in this
case) into ceria improves oxide reducibility, oxygen storage capacity,
and surface and lattice oxygen mobility. These improved features
facilitate the N2O decomposition mechanism steps where ceria is
involved (Steps 4, 5, 6, 7, and 8), that is, ceria doping improves

b1469_Ch-04.indd 228 4/8/2013 12:29:58 PM


b1469 Catalysis by Ceria and Related Materials

Interaction of Nitrogen Oxides with Ceria-Based Materials 229

oxygen transfer from ceria to rhodium and creation of ceria vacant


sites where N2O can be chemisorbed and decomposed.
In Fig. 4.3, H2 reduction profiles of the ceria samples of Fig. 4.1
are shown. Notice the improved reducibility of 0.5 wt% Rh/Ce0.9Pr0.1O2
with regard to 0.5 wt% Rh/CeO2. For 0.5 wt% Rh/Ce0.9Pr0.1O2, the
lowest temperature reduction peak (~110°C), attributed to reduc-
tion of surface Rh-Ce-Pr entities, is more intense, and the intermedi-
ate reduction peak (at 275–300°C), attributed to surface reduction
of ceria, is at a lower temperature.
The catalytic decomposition of N2O on Rh-free ceria needs a
much higher temperature than decomposition on Rh/ceria cata-
lysts (see Fig. 4.1), since Steps 1, 2, 3, 5, and 8 are not in the decom-
position pathway. That is, the N2O decomposition process does not
start with the easy reduction of rhodium sites by N2O but with the
reduction of ceria, which needs a higher temperature. For instance,
in Fig. 4.3 it can be seen that H2 reduction of the La-doped ceria
sample requires a much higher temperature than the surface reduc-
tion of the Rh-containing samples. Hence, samples with very poor
Thermal conductivity detector (TCD)
signal (arbitrary units)

0.5 wt% Rh/Ce0.9Pr0.1O2

0.5 wt% Rh/CeO2

Ce0.9La0.1O2

CeO2 (calcined at 1000°C)

0 100 200 300 400 500 600 700 800 900


Temperature (°C)

Figure 4.3 Temperature-programmed reduction by H2 (H2-TPR) of the ceria


samples of Fig. 4.1.

b1469_Ch-04.indd 229 4/8/2013 12:29:58 PM


b1469 Catalysis by Ceria and Related Materials

230 A. García-García and A. Bueno-López

reducibility, such as ceria calcined at 1000°C (see Fig. 4.3), hardly


decompose N2O (see Fig. 4.1). A relation between N2O decomposi-
tion capacity and H2 reduction of surface ceria was found for a set
of pure and La- or Pr-doped ceria samples with and without noble
metals (Rh, Pd, or Pt), and it was proposed that the rate-limiting
step of the N2O decomposition mechanism is the reduction of cata-
lyst active sites.19
As has been described, the catalytic decomposition of N2O on
Rh/ceria catalysts follows a redox mechanism where not only rho-
dium but also ceria is progressively reduced and oxidized by N2O.
Temperature-programmed reduction by H2 (H2-TPR) is a useful
technique for correlating catalytic decomposition with the reducibil-
ity of the ceria material. The oxidation processes involved in N2O
decomposition have also been studied by in situ Raman spectroscopy
and by X-ray photoelectron spectroscopy (XPS) measurements per-
formed after in situ treatments under different conditions.28 Figure 4.4
shows details of the F2g band, attributed to the fluorite structure of
ceria, measured by in situ Raman spectroscopy performed with Ar

350ºC 150ºC 25ºC


Normalized intensity (arb. units)

Ar

150ºC 25ºC
350ºC

1000 ppm N2O/Ar

400 420 440 460 480 500


-1)
Raman shift (cm

Figure 4.4 F2g band measured by in situ Raman spectroscopy performed with
a 0.5 wt% Rh/CeO2 catalyst in Ar and 1000 ppm N2O/Ar flows at different
temperatures.

b1469_Ch-04.indd 230 4/8/2013 12:29:59 PM


b1469 Catalysis by Ceria and Related Materials

Interaction of Nitrogen Oxides with Ceria-Based Materials 231

and 1000 ppm N2O/Ar flows at different temperatures with a 0.5 wt%
Rh/CeO2 catalyst.
For both gases, the position of the F2g peak decreases with tem-
perature, and this is attributed to the thermal expansion of the unit
cell of ceria due to the partial reduction of Ce4+ cations (0.097 nm)
to larger Ce3+ cations (0.114 nm). Thermal reduction occurs because
of O2 release with temperature, but at 350 ºC, that is, once the tem-
perature is high enough for N2O decomposition (see Fig. 4.1), the
shift of the F2g peak mainly affects the experiment performed in the
inert (Ar) flow. In N2O/Ar, N2O re-oxidizes the thermally reduced
cerium cations. The thermal reduction of ceria in Ar and also the
reoxidation by N2O was confirmed by XPS measurements.28 These
experiments evidence the oxidizing character of N2O towards
reduced ceria, N2O oxidation being an intermediate step on the cata-
lytic decomposition of N2O on Rh/ceria.
DRIFTS is also a powerful tool for studying the interaction of
N2O with ceria-based materials, and as an example, Fig. 4.5 shows
DRIFT spectra recorded at 375°C during in situ experiments per-
formed with two ceria samples of composition 0.25 wt% Rh/50 wt%
Ce0.9Pr0.1O2/γ-Al2O3 and 50 wt% Ce0.9Pr0.1O2/γ-Al2O3 in a 1000 ppm
N2O/He flow.
In general, there are three regions of the infrared (IR) spectra
that merit special attention. Bands in the range 2400–2100 cm−1 can
be assigned to the N–N stretching frequency of adsorbed N2 or
N2O,31–33 bands below 1700 cm−1 can be attributed to oxidized nitro-
gen species (nitrite, nitrate, and nitro compounds), and those
around 3800–3600 cm−1 to hydroxyl groups.34,35 Table 4.1 summa-
rizes the main IR bands identified for a Rh/ceria catalyst in contact
with a N2O/He flow.
Bands of N–N stretching are observed in Fig. 4.5 both in the
Rh-free and Rh-containing samples (2400–2100 cm−1), indicating
that N2O is chemisorbed on ceria (not just on Rh). The formation
of more than a single band occurs due to the presence of different
chemisorption sites, which is an evidence of the heterogeneity of the
ceria surface. Bands between 2400 cm−1 and 2100 cm−1 appear in
ceria samples typically above 200ºC. On γ-Al2O3 and Rh/γ-Al2O3

b1469_Ch-04.indd 231 4/8/2013 12:29:59 PM


b1469 Catalysis by Ceria and Related Materials

232 A. García-García and A. Bueno-López

Sample 50 wt.% Ce0.9Pr0.1O2/γ-Al2O3 without and with 0.25 wt.%Rh 1220


1340
1434
1515
3760 cm-1
Kubelka–Munk units

2400-2100 cm-1

without Rh

with Rh

4000 3600 3200 2800 2400 2000 1600 1200


Wavenumber (cm-1)

Figure 4.5 DRIFT spectra of the samples 0.25 wt% Rh/50 wt% Ce0.9Pr0.1O2/γ-Al2O3
(with rhodium) and 50 wt% Ce0.9Pr0.1O2/γ-Al2O3 (without rhodium) recorded
in situ at 375°C in a gas flow of 1000 ppm N2O/He (baselines recorded at room
temperature in He were subtracted).

Table 4.1 IR assignment of OH groups and surface nitrogen species present on


Rh/ceria catalysts under a N2O/He flow.
IR frequencies (cm−1) Proposed species References
2400–2100 N–N stretching frequency of adsorbed 31–33
N2 or N2O
3800–3600 hydroxyl groups 34, 35
1434 monodentate nitrites 28, 31
1515 and 1220 monodentate nitrates 28, 31
1360–1340 nitro groups 28, 31
1695 and 1230 bridged nitrates 28, 31
1530 and 1300 chelated nitrates 28, 31

samples, for instance, a single band at 2315 cm−1 was observed for
N2O chemisorption above 200°C.28
In Fig. 4.5, the negative band at 3740 cm−1 is attributed to the
depletion of hydroxyl groups on ceria, according to Step 4 of the

b1469_Ch-04.indd 232 4/8/2013 12:29:59 PM


b1469 Catalysis by Ceria and Related Materials

Interaction of Nitrogen Oxides with Ceria-Based Materials 233

mechanism, and the expected release of H2O was confirmed by


online mass spectrometry monitoring of the gas stream.30
Ceria is able to oxidize chemisorbed N2O, and the formation of
bands below 1700 cm−1 occurs due to this oxidation process. Bands
that can be assigned31 to monodentate nitrites (at 1434 cm−1),
monodentate nitrates (1515 cm−1 and 1220 cm−1), and nitro groups
(1340 cm−1) are identified in Fig. 4.5. This oxidation process
depends on the particular features of each experiment and each
sample, and other oxidized nitrogen species can be identified at dif-
ferent temperatures and with other ceria materials.28 It has been
reported for bare ceria (calcined at 600°C)28 that below 200°C the
absorption bands identified below 1700 cm−1 were consistent with
the formation of monodentate nitrates both in N2O/He and N2O/
O2/He streams. Above this temperature, these bands decreased and
there was a corresponding increase of bands at 1695 and 1230 cm−1
attributed to bridged nitrates and at 1530 and 1300 cm−1 to chelated
nitrates. Bridged and bidentate nitrates are more stable than
monodentate nitrates, and therefore they appear at higher tempera-
tures. In addition, bands compatible with the formation of monoden-
tate nitrites (1434 cm−1) and nitro groups (1360 cm−1) remained on
bare ceria between 200° and 350°C. These oxidized nitrogen species
seem not to be reaction intermediates of the N2O decomposition
mechanisms but are just spectators.28
In conclusion the interaction of N2O with ceria materials leads
to the formation of several surface nitrogen groups including N2O
itself and also oxidized species like nitrites and nitrates. In some
conditions (high enough temperature, presence of certain metals
like rhodium, etc.) N2O is decomposed to N2 and O2. With Rh/ceria
catalysts, it has been demonstrated that ceria is actively involved in
the decomposition and the active sites for N2O chemisorption and
decomposition are not only located on rhodium but also on ceria.

4.3 NOx Interaction with Ceria-Based Materials:


Ceria-Catalyzed Oxidation of NO to NO2
Other than at very high temperatures, the catalytic decomposition
of NO into N2 and O2 is thermodynamically favorable36,37 but the

b1469_Ch-04.indd 233 4/8/2013 12:29:59 PM


b1469 Catalysis by Ceria and Related Materials

234 A. García-García and A. Bueno-López

decomposition rate is so slow that NO is kinetically stable. Catalysts


could accelerate the decomposition rate but so far no catalyst with a
sufficiently high activity and stability for commercial application is
known.36
Usually catalysts for NOx removal must deal with NO + O2
mixtures, as in the de-NOx process for diesel vehicles or in coal-
combustion power plants. In contrast to N2O, NO’s interaction with
ceria-based materials (and many other catalysts) in O2-rich streams
does not lead to the decomposition of the molecule but to oxidation
to NO2. This process involves NOx chemisorption and the formation
of several surface nitrogen species, oxidation of these surface spe-
cies, and desorption of NO2. In addition, NO2 can be re-adsorbed on
the catalyst after release. All these processes are discussed in this
section.
Figure 4.6 shows the NOx chemisorption and NO2 formation
curves obtained with two ceria samples in contact with a gas flow of
500 ppm NO + 5% O2 while heating at 10°C/min.
Low-temperature calcined CeO2 (at 500°C in the example of
Fig. 4.6) chemisorbs NOx above approximately 250°C, and the
chemisorbed NOx species are stable in the NOx-containing gas
stream since negative values are not observed in the NOx chemisorp-
tion profiles. The presence of NOx in the gas mixture stabilizes
chemisorbed NOx species on the ceria surface, otherwise NOx would
be released. For instance, bulk cerium nitrate can be totally decom-
posed to cerium oxide in air or in inert gas at ca. 500°C.38
In addition to NOx chemisorption, low-temperature calcined
CeO2 (at 500°C in the example of Fig. 4.6) is also able to oxidize
NO to NO2 from 300°C (Fig. 4.6(b)). The NO2 profile increases
with temperature until the thermodynamic NO oxidation equilib-
rium is reached, and then decreases at higher temperatures follow-
ing the thermodynamics. The highest NO2 concentration is
reached at ca. 450°C, and most ceria catalysts typically reach the
maximum NO2 concentration around this temperature or at higher
values. Pt catalysts, for instance, which are much more active than
ceria for NO oxidation, reach the maximum NO2 production at ca.
250°C.

b1469_Ch-04.indd 234 4/8/2013 12:29:59 PM


b1469 Catalysis by Ceria and Related Materials

Interaction of Nitrogen Oxides with Ceria-Based Materials 235

20

16 CeO2 (calcined at 500ºC)


NOx chemisorbed (%)

12

4 CeO2 (calcined at 1000ºC)

0
225 325 425 525 625
Temperature (ºC)
(a)
40
Thermodynamic equilibrium
NO + 1/2O2 NO2
30
CeO2 (calcined at 500ºC)
NO2 (%)

20

10
CeO2 (calcined at 1000ºC)
Empty reactor

0
225 325 425 525 625
Temperature (ºC)
(b)

Figure 4.6 Interaction of a 500 ppm NO + 5% O2 gas flow (30,000 h−1) at atmos-
pheric pressure with powdered ceria samples calcined at different temperatures.
Experiments were performed in a fixed-bed reactor heated at 10°C/min and cou-
pled to NDIR-UV specific gas analyzers. (a) NOx chemisorbed (%) = 100·([NOx]in-
[NOx]out)/ [NOx]in and (b) NO2 (%) = 100·[NO2]out/([NO]out + [NO2]out).

b1469_Ch-04.indd 235 4/8/2013 12:29:59 PM


b1469 Catalysis by Ceria and Related Materials

236 A. García-García and A. Bueno-López

As will be described below, the NOx chemisorption and NO2


production processes occurring on ceria are related to each other,
and chemisorbed NOx species are reaction intermediates of the
ceria-catalyzed NO oxidation mechanism.
Both NOx chemisorption and NO oxidation processes are
affected by ceria features. One important parameter for explaining
the amount of NOx chemisorbed and the extent of NO oxidation is
the BET surface area of the ceria material, and typically for ceria
samples of similar composition the higher the surface area, the
higher the NOx chemisorption and NO oxidation capacity. This is
one of the arguments, as seen in Fig. 4.6, to explain why 1000°C
calcined ceria (BET=2 m2/g) has a poor NO2 production capacity
with regard to the counterpart ceria calcined at 500°C (BET=65 m2/g),
but this is not the only reason. The nature of the ceria surface also
significantly affects its interaction with NOx, and the surface proper-
ties of ceria are in some cases even more important than the BET
surface area especially for doped cerias.
A representative example of the important role of surface com-
position for the interaction of ceria materials with NOx was a study27
about the effect of the amount of cerium and zirconium on the
adsorption of NOx on CexZr1-xO2 mixed oxides. This study was car-
ried out with a set of CexZr1-xO2 samples with x = 1, 0.76, 0.56, 0.36,
0.16, and 0, and it concluded that their catalytic activity for NO oxi-
dation to NO2 mainly depends on the cerium molar fraction. A lin-
ear relation between NO2 production capacity and cerium content
was found for these ceria–zirconia mixed oxides. However, bare
CeO2 did not follow the same trend as the mixed oxides, and its cata-
lytic activity for NO oxidation was lower than expected considering
its cerium content. This was mainly attributed to the availability of
reactive oxygen needed for NO oxidation, which is enhanced in
doped cerias.
In brief, the ceria-catalyzed oxidation of NO to NO2 involves NO
chemisorption on the catalyst, oxidation of the nitrogen surface spe-
cies created on NO chemisorption, and NO2 desorption. In order to
analyze what occurs at the surface of ceria in contact with NO + O2
mixtures at temperatures around 350ºC, a set of CexZr1−xO2 catalysts

b1469_Ch-04.indd 236 4/8/2013 12:29:59 PM


b1469 Catalysis by Ceria and Related Materials

Interaction of Nitrogen Oxides with Ceria-Based Materials 237

Table 4.2 IR assignment of OH groups and surface nitrogen species present on


CexZr1-xO2 catalysts (with x from 0 to 1) in NOx flows at temperatures around
350 °C. (Note that IR frequencies and relative intensities are sensitive to the Ce/Zr
surface ratio and to the preparation method.)
IR frequencies (cm−1) Proposed species References
3670 and 3660 hydroxyl groups on monoclinic 39–42
zirconia
3730 and 3647 hydroxyl groups on tetragonal zirconia 39–42
3715–3500 hydroxyl groups on ceria 39–42
3760–3700 hydroxyl groups on ceria–zirconia 27, 31, 43
mixed oxides
1190 bidentate nitrites on zirconia 27, 43–45
1184 bidentate nitrites on ceria 27, 43–45
1200–1105 bidentate nitrites on ceria–zirconia 27, 42–44
mixed oxides
1620, 1230, 1000 bridged nitrates on zirconia 31
1583, 1242, 1030 chelated nitrates on zirconia 42
1550–1450, 1280, 1000 monodentate nitrates on zirconia 42
1595, 1210, 1000 bridged nitrates on ceria 27, 42–45
1585–1545, 1265–1225, chelated nitrates on ceria 27, 42–45
1030–1000
1540–1500, 1275, 1030 monodentate nitrates on ceria 27, 42–45
1600 and above, 1230, bridged nitrates on ceria–zirconia 42
1000 mixed oxides
1600–1545, 1260–1230, chelated nitrates on ceria–zirconia 42
1030–1000 mixed oxides
1545–1500, 1270, 1030 monodentate nitrates on ceria–zirconia 42
mixed oxides

was selected and in situ DRIFTS experiments and NOx adsorption


plus temperature-programmed desorption (TPD) experiments were
performed. Table 4.2 shows some of the relevant IR frequencies
reported in different studies on the interaction of NOx with ceria
and ceria–zirconia materials.
It has been reported31 that as the surface coverage of nitrogen
species increases, discrete frequency shifts to higher wavenumbers

b1469_Ch-04.indd 237 4/8/2013 12:29:59 PM


b1469 Catalysis by Ceria and Related Materials

238 A. García-García and A. Bueno-López

were observed in some bands. Usually this phenomenon is explained


by repulsive interactions between the adsorbate molecules when
their local concentrations are high.
From the in situ DRIFTS and NOx adsorption plus TPD experi-
ments performed with CexZr1-xO2 catalysts, the following reactions
were proposed to occur during the ceria-catalyzed NO oxidation to
NO2:

NO + 2Ce4+−OH− → Ce4+−NO2− + Ce3+−… + H2O Step 9


Ce3+−… + NO → Ce4+ −NO− Step 10
Ce3+−… + NO2 → Ce4+−NO2– Step 11
Ce4+−NO− + Ce4+−O* → Ce4+−NO2−+ Ce3+−… Step 12
Ce4+−NO2− + Ce4+−O* → Ce4+−NO3− + Ce3+−… Step 13
NO2 + 2Ce4+−OH− → Ce4+− NO3− + Ce3+−… + H2O Step 14
2NO2 → N2O4 → NO+−NO3− Step 15
Ce3+−… + NO2 → Ce4+−O* + NO Step 16

In these reactions, “Ce3+−…” represents an oxygen vacancy linked to


a Ce3+ cation, Ce4+–OH− a hydroxyl group on a Ce4+ cation, and
Ce4+−O* is an active oxygen on ceria. In ceria–zirconia mixed
oxides, the Zr4+−OH− groups could also react with NO, as in Step 9,
but OH− on Ce4+ are expected to be more reactive than OH− on Zr4+.
Figure 4.7 shows the main reactions leading to the catalytic oxi-
dation of NO to NO2 on ceria-based materials.
In the early stages of the NO + O2 interaction with ceria materi-
als, the experiments identified chelated and bridged nitrites (NO2−),
which can be formed by reaction with OH groups or on NO
chemisorption on ceria vacants (Steps 9 and 10, respectively). When
NO2 is formed by oxidation of NO, NO2 re-adsorption can yield
nitrites (Step 11). The existence of anionic NO− species, as sug-
gested by Step 10, has not been proven because their IR absorption
closely matches that of chelated and bridged nitrites. However, con-
sidering the strong oxidizing properties of ceria-based materials, it
is possible that NO− anions exist as transient species in the electron
and oxygen transfer processes involved in the formation of nitrites
from NO.

b1469_Ch-04.indd 238 4/8/2013 12:29:59 PM


b1469 Catalysis by Ceria and Related Materials

Interaction of Nitrogen Oxides with Ceria-Based Materials 239

NO
Re-adsorbed NO2 Gas phase NO2

Nitrites Nitrates
OH

O
CeO2 CeO2-x CeO2
Oxygen
vacancies

O2

Figure 4.7 Main reactions leading to the catalytic oxidation of NO to NO2 on


ceria.

With longer reaction times, the experiments identified nitrates


in a variety of structures and configurations, including bridged,
chelated, and monodentate nitrates (NO3−). These species were
formed at the expense of nitrites, and Step 13 describes the oxida-
tion of nitrites to nitrates by ceria oxygen. An alternative pathway
leading to nitrates from the reaction of ceria–zirconia with NO + O2
requires NO2 gas (formed via NO oxidation) as an intermediate
(Step 14). A third reaction pathway that would also lead to the for-
mation of surface nitrates, and does not require electron transfer
from/to the catalyst surface, uses molecularly adsorbed NO2 in the
form of a N2O4 dimer (Step 15). The self-disproportionation of the
dimer, which has been shown to occur on terrace sites of TiO2 or
defect-free MgO,31 is known to create a local pair of charged adsorb-
ates formed by electron transfer from a NO2 adsorbate to another
and an acid-base interaction with the surface.
The sequence of Steps 9–15 leads to the depletion of OH
groups and active oxygen (Ce4+−O*), that is, there is a net consump-
tion of oxidizing surface species. The steady-state oxidation of NO
to NO2 requires that some of these oxidizing species are restored.

b1469_Ch-04.indd 239 4/8/2013 12:29:59 PM


b1469 Catalysis by Ceria and Related Materials

240 A. García-García and A. Bueno-López

In NO + O2 mixtures, active oxygen can be created on ceria by O2


chemisorption and dissociation, and also by NO2 re-adsorption
(Step 16) but the latter process does not contribute to net NO2 pro-
duction since NO2 is consumed.
The relative rate of these steps depends on the particular com-
position of the ceria material. Figure 4.8 shows the transformation
over time of surface nitrites and nitrates created on CeO2, ZrO2, and
Ce0.76Zr0.24O2 samples during in situ DRIFTS experiments performed
at 350°C with a NO + O2 stream.
As described, nitrites appear first in all samples, and after a few
minutes, nitrites on pure and 24% zirconium-doped ceria are pro-
gressively oxidized to nitrates while those on zirconia are oxidized
much more slowly. In a more complete study performed with
CexZr1-xO2 mixed oxides of different compositions,27 it was con-
cluded that the formation of nitrates occurs at much higher rates for
the intermediate and Ce-rich compositions (x = 1, 0.76, 0.56, and
0.36) than for the Zr-rich ones (x = 0 and 0.16). This rate was faster
for Ce0.76Zr0.24O2 and Ce0.56Zr0.44O2 than for pure CeO2, confirming
Absorbance of the 1180 cm-1 band/arbitrary units
Area in the 1460-1650 cm-1 IR region/arbitrary units

ZrO2
Surface nitrites population
Surface nitrates population

Ce0.76Zr0.24O2

CeO2

ZrO2
CeO2

Ce0.76Zr0.24O2

0 10 20 30 40 50 60 70
Time (min)

Figure 4.8 Transformation over time of surface nitrites (open symbols) and
nitrates (solid symbols) on CeO2, ZrO2, and Ce0.76Zr0.24O2 during in situ DRIFTS
experiments performed at 350°C with 2000 ppm NO + 5% O2.

b1469_Ch-04.indd 240 4/8/2013 12:29:59 PM


b1469 Catalysis by Ceria and Related Materials

Interaction of Nitrogen Oxides with Ceria-Based Materials 241

that ceria doping by zirconium improves the oxidation capacity of


the mixed oxide. Conversely, ZrO2 was characterized by the highest
volume of nitrites (or nitrosyl anion) and the lowest volume of
nitrates. The trends observed in the formation rate of surface
nitrates reflect somewhat the activities measured during NO oxida-
tion tests. On the other hand, the participation of OH groups in the
formation of ad-NOx species was also confirmed, and the most reac-
tive OH groups are those linked to Ce atoms. In practice, these reac-
tive OH groups are quickly depleted and this should in turn result
in the formation of new sorption sites.
The rate of the catalyzed oxidation of NO to NO2 at around
350 °C followed roughly the same trend observed for the production
of nitrates, but NO2 production is also affected by the adsorption
strength of the nitrogen species on the ceria surface. On Zr-rich
surfaces, the thermostability of NOx ad-species is high. Zr-rich
cerium–zirconium mixed oxides contain strong NOx adsorption
sites, and the nitrite to nitrate conversion rate is slow on these sur-
faces, which explains their poor catalytic activity for NO2 production
in comparison with cerium-rich formulations. In other words, the
ceria materials catalyzed oxidation of NO to NO2 in NO + O2 mix-
tures not only depends on the NO/nitrites oxidation capacity of the
catalyst used, but also on the adsorption strength of the nitrates
under NO oxidation conditions.27 This can be seen from Fig. 4.9,
which shows NO2 profiles obtained in TPD experiments performed
in a He flow after NOx chemisorption at 350°C with a mixture of
2000 ppm NO + 5%O2. It can be seen that NO2 on ZrO2 needs a
higher temperature to be released than NO2 on CeO2 or Ce0.76Zr0.24O2.
The bimodal shape of the mixed oxide curve evidences the pres-
ence of NOx adsorption sites of different strengths, and the lowest
area of the ceria curve is attributed to the previous desorption of
loosely bound nitrogen groups during the cleaning step performed
at 350°C in He before TPD.
All these steps for the catalytic oxidation of NO to NO2 (chem-
isorption plus surface oxidation plus desorption) are greatly
dependent on the number and type of sites present on the ceria-
material surface and variables affecting these sites affect the

b1469_Ch-04.indd 241 4/8/2013 12:30:00 PM


b1469 Catalysis by Ceria and Related Materials

242 A. García-García and A. Bueno-López

ZrO2
NO2 concentration (arbitrary untis)

Ce0.76Zr0.24O2

CeO2

350 375 400 425 450 475 500 525 550


Temperature (°C)

Figure 4.9 Temperature-programmed desorption (TPD) of NO2 in helium after


NO + O2 adsorption at 350°C on CeO2, ZrO2, and Ce0.76Zr0.24O2. (A one-hour
cleaning step was performed at 350°C in He after NOx chemisorption before TPD
during which the loosely bound nitrogen species were desorbed.)

interaction with NOx. One important variable, for instance, is the


calcination temperature, as shown by Fig. 4.6. The catalytic activity
for NO oxidation of the mixed oxide Ce0.76Zr0.24O2 calcined at dif-
ferent temperatures from 500 to 1000ºC was studied,44 concluding
that the calcination temperature decreases the BET area and also
the Ce/Zr atomic surface ratio (determined by XPS) and both
factors affected the rate of nitrite to nitrate transformation and the
further desorption of NO2. It is usually difficult to separate the
effect of surface area from that of other variables relating to sur-
face composition, but further studies performed with cerium–
zirconium mixed oxides of similar composition but prepared by
different methods46,47 demonstrated that the surface concentration
of ceria is a critical parameter for explaining the catalytic oxidation
of NO to NO2. As a general trend, it can be concluded that ceria-
rich surfaces are more active than those with a high zirconium
concentration.

b1469_Ch-04.indd 242 4/8/2013 12:30:00 PM


b1469 Catalysis by Ceria and Related Materials

Interaction of Nitrogen Oxides with Ceria-Based Materials 243

4.4 Conclusions
In this chapter, the interaction of nitrogen oxides (N2O, NO, and
NO2) with ceria-based materials has been discussed, with a focus on
two catalyzed reactions of practical relevance: the catalytic decom-
position of N2O on Rh/ceria catalysts and the ceria-catalyzed oxida-
tion of NO to NO2.
The interaction of N2O with ceria materials leads to the formation
of several surface nitrogen groups, including N2O itself but also oxi-
dized species like nitrites and nitrates due to the highly oxidizing
character of ceria. In some conditions (a high enough temperature,
the presence of certain metals like rhodium, etc.) N2O is decomposed
to N2 and O2. For Rh/ceria catalysts, it has been demonstrated that
ceria is actively involved in the decomposition and the active sites for
N2O chemisorption and decomposition are not only located on rho-
dium but also on ceria.
In contrast to N2O, the interaction of NOx (NO + NO2) with
ceria-based materials does not yield N2 + O2 and the main reaction
is the oxidation of NO to NO2. This oxidation process involves NO
chemisorption on the catalyst, oxidation of the nitrogen surface spe-
cies created, and NO2 desorption. The surface nitrogen species
identified on the ceria surface after NOx chemisorption are nitrites,
nitrates, and nitro groups. These species are progressively converted
with time to more oxidized nitrogen groups. After these oxidation
processes NO2 is released. Ceria oxygen is used in these surface oxi-
dation steps, but it is not only the oxidizing character of the ceria
material that affects the net oxidation of NO to NO2. NO2 produc-
tion is also affected by the adsorption strength of the nitrogen
species on the ceria surface. For ceria–zirconia mixed oxides, for
instance, surface nitrogen groups are attached more strongly to Zr
sites than to Ce sites.
The commonality of the interaction of N2O and NOx with ceria-
based materials is the important role of the properties of ceria,
which are affected, as is well known, by variables like the calcination
temperature, dopants, etc. A ceria-based material with “good” prop-
erties (easy reducibility, high oxygen storage capacity, high surface

b1469_Ch-04.indd 243 4/8/2013 12:30:00 PM


b1469 Catalysis by Ceria and Related Materials

244 A. García-García and A. Bueno-López

and lattice oxygen mobility, etc.) promotes some of the processes. In


the catalytic decomposition of N2O on Rh/ceria, ceria with good
properties improves the catalytic decomposition, since it facilitates
the transfer of oxygen from ceria to rhodium and the creation of
active sites on ceria suitable for N2O chemisorption and decomposi-
tion. In the catalytic oxidation of NO to NO2, ceria with good prop-
erties enhances the oxidation of surface nitrogen species, which is
an intermediate step in oxidation.

References
1. Greenwood, N.N., Earnshaw, A., Chemistry of the Elements, 2nd edition,
Elsevier, Oxford, (1997).
2. Kapteijn, F., Rodríguez-Mirasol, J., Moulijn, J.A., Appl. Catal. B, 9
(1996) 25–64.
3. Pérez-Ramírez, J., Kapteijn, F., Schoffel, K., Moulijn, J.A., Appl. Catal.
B, 44 (2003) 117–151.
4. Lee, S.J., Ryu, I.S., Kim, B.M., Moon, S.H., Int. J. Greenh. Gas Con., 5
(2011) 167–176.
5. Gandhi, H.S., Graham, G.W., McCabe, R., J. Catal., 216 (2003)
433–442.
6. Garrigós-Pastor, G., Parres-Escaplez, S., Bueno-López, A., Illán-Gómez,
M.J., Salinas-Martínez de Lecea, C., Appl. Catal. A, 354 (2009) 63–71.
7. Bueno-López, A., Lozano-Castello, D., Such-Basáñez, I., García-Cortés,
J.M., Illán-Gómez, M.J., Salinas-Martínez de Lecea, C., Appl. Catal. B,
58 (2005) 1–7.
8. Shimizu, A., Tanaka, K., Fujimori. M., Chemosphere, 2 (2000) 425–434.
9. Granger, P., Esteves, P., Kieger, S., Navascues, L., Leclercq, G., Appl.
Catal. B, 62 (2006) 236–243.
10. Alini, S., Basile, F., Blasioli, S., Rinaldi, C., Vaccari, A., Appl. Catal. B,
70 (2007) 323–329.
11. Pérez-Ramírez, J., Santiago, M., Chem. Commun., 6 (2007) 619–621.
12. Giecko, G., Borowiecki, T., Gac, W., Kruk, J., Catal. Today, 137 (2008)
403–409.
13. Ruszak, M., Inger, M., Witkowski, S., Wilk, M., Kotarba, A., Sokja, Z.,
Catal. Letters, 126 (2008) 72–77.

b1469_Ch-04.indd 244 4/8/2013 12:30:00 PM


b1469 Catalysis by Ceria and Related Materials

Interaction of Nitrogen Oxides with Ceria-Based Materials 245

14. Bueno-López, A., Such-Basáñez, I., Salinas-Martínez de Lecea, C.,


J. Catal., 244 (2006) 102–112.
15. Centi, G., Perathoner, S., Rak, Z.S., Appl. Catal. B, 41 (2003)
143–155.
16. Suárez, S., Yates, M., Gil Llambías, F.J., Martín, J.A., Avila, P., Blanco,
J., Stud. Surf. Sci. Catal., 143 (2000) 111–119.
17. Imamura, S., Tadani, J., Saito, Y., Okamoto, Y., Jindai, H., Kaito, C.,
Appl. Catal. A, 201 (2000) 121–127.
18. Rico-Pérez, V., Parres-Esclapez, S., Illán-Gómez, M.J., Salinas-Martínez
de Lecea, C., Bueno-López, A., Appl. Catal. B, 107 (2011) 18–25.
19. Parres-Esclapez, S., Illán-Gómez, M.J., Salinas-Martínez de Lecea, C.,
Bueno-López, A. Appl. Catal. B, 96 (2010) 370–378.
20. Fritz, A., Pitchon, V., Appl. Catal. B, 13 (1997) 1–25.
21. Kaspar, J., Fornasiero, P., Graziani, M., Catal. Today, 50 (1999)
285–298.
22. Setiabudi, A., Chen, J., Mul, G., Makkee, M., Moulijn, J.A., Appl. Catal.
B, 51 (2004) 9–19.
23. Aneggi, E., de Leitenburg, C., Dolcetti, G., Trovarelli, A., Catal. Today,
114 (2006) 40–47.
24. Krishna, K., Bueno-López, A., Makkee, M., Moulijn, J.A., Appl. Catal.
B, 75 (2007) 189–200.
25. Krishna, K., Bueno-López, A., Makkee, M., Moulijn, J.A., Appl. Catal.
B, 75 (2007) 201–209.
26. Krishna, K., Bueno-López, A., Makkee, M., Moulijn, J.A., Appl. Catal.
B, 75 (2007) 210–220.
27. Atribak, I., Azambre, B., Bueno-López, A., García-García, A., Appl.
Catal. B, 92 (2009) 126–137.
28. Parres-Esclapez, S., Such-Basañez, I., Illán-Gómez, M.J., Salinas-
Martínez de Lecea, C., Bueno-López, A., J. Catal., 276 (2010)
390–401.
29. Cunningham, J., Hickey, J.N., Cataluna, R., Conesa, J.C., Soria, J.,
Martinez-Arias, A. Stud. Surf. Sci. Catal., 101 (1996) 681–690.
30. Parres-Esclapez, S., Illán-Gómez, M.J., Salinas-Martínez de Lecea, C.,
Bueno-López, A., Int. J. Greenh. Gas Con. 11(2012), 251–261.
31. Hadjiivanov, K.I., Catal. Rev. Sci. Eng., 42 (2000) 71–144.

b1469_Ch-04.indd 245 4/8/2013 12:30:00 PM


b1469 Catalysis by Ceria and Related Materials

246 A. García-García and A. Bueno-López

32. Oktar, N., Mitome, J., Holmgreen, E.M., Ozkan, U.S., J. Mol. Catal., A.
259 (2006) 171–182.
33. Wang, Y., Lei, Z., Chen, B., Guo, Q., Liu, N., Appl. Surf. Sci., 256 (2010)
4042–4047.
34. Djinović, P., Batista, J., Pintar, A. Catal. Today, 147 (2009) 191–197.
35. Menacherry, P.V., Haller, G.L., Catal. Letters, 44 (1997) 135–144.
36. Kustova, M.Yu., Rasmussen, S.B., Kustov, A.L., Christensen, C.H., Appl.
Catal. B, 67 (2006) 60–67.
37. Parvulescu, V.I., Granger, P., Delmon, B., Catal. Today, 46 (1998)
233–316.
38. Miguel-García, I., Parres-Esclapez, S., Lozano-Castelló, D., Bueno-
López, A. Catal. Commun., 11 (2010) 848–852.
39. Binet, C., Daturi, M., Lavalley, J.C. Catal. Today, 50 (1999) 207–225.
40. Daturi, M., Finocchio, E., Binet, C., Lavalley, J.C., Fally, F., Perrichon,
V., J. Phys. Chem. B, 103 (1999) 4884–4891.
41. Graf, P.O., de Vlieger, D.J.M., Mohet, B.L., Lefferts, L., J. Catal., 262
(2009) 181–187.
42. Azambre, B., Atribak, I., Bueno-López, A., García-García, A., J. Phys.
Chem. C, 114 (2010) 13300–13312.
43. Azambre, B., Zenboury, L., Koch, A., Weber, J.A., J. Phys. Chem. C, 113
(2009) 13287–13299.
44. Atribak, I., Azambre, B., Bueno-López, A., García-García, A., Top.
Catal., 52 (2009) 2092–2096.
45. Martinez-Arias, A., Soria, J., Conesa, J.C., Seoane, X.L., Arcoya, A.,
Cataluna, R.J., J. Chem. Soc. Faraday T., 91 (1995) 1679–1687.
46. Azambre, B., Atribak, I., Bueno-López, A., García-García, A., Phys.
Chem. Chem. Phys., 12 (2010) 13770–13779.
47. Atribak, I., Guillén-Hurtado, N., Bueno-López, A., García-García, A.,
Appl. Surf. Sci., 256 (2010) 7706–7712.

b1469_Ch-04.indd 246 4/8/2013 12:30:00 PM


b1469 Catalysis by Ceria and Related Materials

CHAPTER 5

ATOMISTIC MODELLING OF CERIA


NANOSTRUCTURES: INTRODUCING
STRUCTURAL COMPLEXITY
Dean C. Sayle and Thi X. T. Sayle
Department of Engineering and Applied Science,
Cranfield University,
Defence Academy of the United Kingdom,
Shrivenham, UK, SN6 8LA

5.1 Methodology
Atomistic computer simulation is an intuitive approach in that it
involves generating the (atomistic) structure and then calculating a
particular property or simulating a process using this structure.
Many texts have been devoted to describing such an approach with
various levels of detail, for example, Catlow et al.1 Accordingly, only
the salient details are provided here. For example, to explore the
oxidative catalysis of ceria one can envisage determining how easy it
is to remove an oxygen atom from the surface of ceria. Accordingly,
one first generates an atomistic model — in this case an array of
atoms to represent the crystal structure of ceria and the particular
surface. One then calculates the energy of this structure compared
to the energy of the same structure after having removed oxygen
from the surface. The study can then be extended to a range of
surfaces to help predict which surfaces are likely to be catalytically
reactive.

247

b1469_Ch-05.indd 247 4/8/2013 12:31:04 PM


b1469 Catalysis by Ceria and Related Materials

248 D. C. Sayle and T. X. T. Sayle

However, before any simulations can be undertaken one needs to


describe the interactions between, for example, the Ce and O atoms,
coined the ‘forcefield’ or ‘potential model’, and also a way of moving
the atoms into a low-energy configuration (theoretical method).

5.1.1 Potential models


There is a wealth of functional forms to describe the interactions
between ions, which generally vary according to the class of mate-
rial: metal, semiconductor, insulator, ionic or covalent. Ceria is an
ionic material and is well suited to being described using formal
charges, which are balanced by repulsive interactions, the latter
associated with neighbouring electron clouds. A popular functional
form used to describe the interaction between ions i and j, is the
Born model of an ionic solid, which can be written:
Ê - rij ˆ
qi q j ÁË ρ ˜¯
Eij = Â + Â Ae
rij

where Eij is the interaction energy of ions i and j, rij the interatomic
distance between ions i and j, qi the charge of ion i and A and ρ are
variable parameters. The first term of the equation is the classical cou-
lombic term, and the second term is the short-range repulsion
between ions. The accuracy and reliability of all atomistic simulation
rests critically upon the potential models used to describe the inter-
actions between the component ions comprising the material.
Accordingly, careful consideration and choice of the variable param-
eters, in this case, A and ρ, must be exercised.

5.1.2 Theoretical methods


Theoretical methods may take one of two forms. The first, such as
molecular dynamics (MD), allows atoms to move in time into differ-
ent (possibly low-energy) positions. The other type is stochastic in
that mathematical algorithms are used to determine low-energy
(‘stable’) configurations for the atoms.

b1469_Ch-05.indd 248 4/8/2013 12:31:04 PM


b1469 Catalysis by Ceria and Related Materials

Atomistic Modelling of Ceria Nanostructures: Introducing Structural Complexity 249

5.1.2.1 Molecular dynamics


In molecular dynamics the forces acting upon the atoms are calcu-
lated and used to determine the acceleration, velocity and positions
of all the atoms (according to Newtonian mechanics). Accordingly,
one is able to simulate the system as it evolves as a function of time.2

5.1.2.2 Energy minimisation


With energy minimisation a trial (model) structure is constructed,
using, for example, symmetry operators, and the atoms are moved
to reduce the force associated with the atomistic structure of the
model. The procedure is normally iterative and results in a low-
energy ‘equilibrium’ configuration. The methods for minimising
the energy are diverse and include, for example, simple gradient
optimisation of the force-atom distance functions,3 Monte Carlo4
and even genetic algorithms,5 which exploit evolutionary processes.

5.2 Introducing Microstructure


The microstructure plays a pivotal role with respect to the chemical
properties of a material. Accordingly, if one wishes to use atomistic
models to explore the reactivity, then the microstructure must be
captured within the atomistic model including: the morphology and
surfaces exposed, intrinsic and extrinsic point defects, dislocations
and grain boundaries. In the following we use selected examples to
describe how such features are introduced.

5.2.1 The pristine parent material


Many catalytically important oxide materials, such as ceria, are crys-
talline at (catalytic) operating temperatures and therefore a model
of the crystal structure can be generated using symmetry operators.
These symmetry operators, which have been introduced into many
simulation codes (such as GULP), can be used to ‘build’ the atom-
istic model using pertinent crystallographic data. This includes the

b1469_Ch-05.indd 249 4/8/2013 12:31:04 PM


b1469 Catalysis by Ceria and Related Materials

250 D. C. Sayle and T. X. T. Sayle

space symmetry, lattice parameter and atom coordinates, which can


be easily extracted experimentally using, for example, x-ray diffrac-
tion. Because a crystal structure has long-range order it means that
periodic boundary conditions can be used to effectively replicate
the material infinitely in space and therefore one does not have to
simulate in excess of 1023 atoms, indicative of the real material.

5.2.2 Point defects: oxygen vacancies


Equipped with the perfect (parent) material, one can then start to
introduce structural complexity. Perhaps the simplest form is that of
a vacancy. An atom is removed from the lattice and then a particular
simulation strategy, such as energy minimisation or molecular
dynamics, can be used to move (neighbouring) atoms into a low-
energy configuration. The energy difference between the perfect
material and the system, which includes a vacancy, is the vacancy
formation energy. This deceptively simple calculation has important
implications — especially for catalysis — as it provides an insight
into how easy (energetically) it is for the material to participate in
an oxidation reaction by providing a labile source of oxygen.6

5.2.3 Oxygen transport


Once an oxygen vacancy has been created the material needs to
replenish the supply of oxygen. The transport properties of oxygen
in a ceria lattice are key, as facile mobility of oxygen will enable the
material to replenish oxygen lost from the surface by oxygen within
the bulk regions. In particular, simulations have revealed that the
lowest-energy pathway for oxygen to move within a ceria lattice is
from a lattice site to a neighbouring vacancy;6 a schematic is shown
in Fig. 5.1. Ionic transport is vacancy driven, which is important
because knowledge of the mechanism will enable one to envisage
factors that might influence such a process with an aim of increasing
or inhibiting such processes. Clearly, ionic conductivity can be meas-
ured experimentally. However, simulation offers the possibility of
calculating the mobility in the pristine material — deconvoluted

b1469_Ch-05.indd 250 4/8/2013 12:31:04 PM


b1469 Catalysis by Ceria and Related Materials

Atomistic Modelling of Ceria Nanostructures: Introducing Structural Complexity 251

Vacancy Oxygen

Figure 5.1 Vacancy mechanism underpinning oxygen transport in ceria.

with respect to other (structural) features, which cannot yet be


achieved experimentally. Accordingly, simulation offers a valuable
complementary technique.
The removal of oxygen will necessarily be associated with the
reduction of (two) cerium ions — from Ce(IV) to Ce(III) to restore
charge neutrality. The presence of these ions is likely to influence
the oxygen vacancy formation energy and oxygen transport. Indeed,
calculations have confirmed that oxygen vacancies are strongly
bound to Ce(III) species; the larger size of the Ce(III) ion, com-
pared to Ce(IV), was also calculated to have a large impact upon
ionic transport. Clearly, one is not just limited to Ce and therefore a
range of extrinsic dopants have been explored to identify cations
that may help facilitate facile oxygen transport. In particular,
Balducci and co-workers calculated oxygen mobility as a function of
dopant ion radius, which revealed that oxygen transport can be
optimised by doping the ceria with gadolinia7 — a prediction pro-
foundly important also to the field of solid oxide fuel cells.
Although simulation has the capability to deconvolute particular
microstructural features, which a capability experiment endeavours
to achieve, the results can be naïvely incorrect when compared with
experiments because the desirable (in this case oxidative) process
will likely be influenced strongly by other factors. This may include,
for example, complex microstructures, which are not included in
the atomistic model of the pristine material. Accordingly, simulation
data needs to be considered carefully in context with experimental
findings.

b1469_Ch-05.indd 251 4/8/2013 12:31:04 PM


b1469 Catalysis by Ceria and Related Materials

252 D. C. Sayle and T. X. T. Sayle

5.2.4 Surfaces
For oxidative catalysis using ceria, oxygen is necessarily extracted
from the surface of the material. Similar to constructing models for
the bulk material, symmetry operators can be used to construct
atomistic models of surfaces. Generally one needs simply to specify
the Miller index of the particular surface in the code, which then
constructs an atomistic model; an unrelaxed CeO2(310) surface is
shown in Fig. 5.2.
Two strategies for simulating surfaces at the atomistic level have
been considered. The simplest approach uses 3D periodic boundary
conditions and includes a large void to represent the free surface.8
Alternatively, the surface can be considered explicitly using 2D peri-
odic boundary conditions and requires special consideration of the
electrostatics.9
The atomistic structure of a surface is not generally a simple
termination of the bulk material; the surface atoms undergo relaxa-
tion and rumpling. This is hardly surprising as the environment,
such as the coordination number, of a surface atom is profoundly
different from that of an equivalent atom in the bulk. The surface
relaxation can be captured using energy minimisation or molecular
dynamics, which direct the (surface) atoms into low-energy

SURFACE

Figure 5.2 Atom positions comprising an unrelaxed CeO2(310) surface. Cerium


is coloured white and oxygen is dark.

b1469_Ch-05.indd 252 4/8/2013 12:31:04 PM


b1469 Catalysis by Ceria and Related Materials

Atomistic Modelling of Ceria Nanostructures: Introducing Structural Complexity 253

positions. Such surface relaxation and rumpling will influence the


surface reactivity and hence catalytic properties and therefore
must be captured within the atomistic model. Surface sensitive
experimental techniques, such as atomic force microscopy (AFM),
scanning tunnelling microscopy (STM) and low-energy electron dif-
fraction (LEED), have confirmed predictions made using atomistic
simulation.
Calculations performed on ceria have revealed6 that the (111)
surface is energetically the most stable surface followed by (110)
and (310). The (100) surface, which is dipolar and therefore inher-
ently unstable,10 has also been simulated and, via a structural rear-
rangement of the surface atoms, it was possible to quench the
surface dipole.11 The calculations also predicted the (100) surface to
be relatively stable and likely therefore to be present in a real mate-
rial.10 Indeed, catalytically reactive cuboidal ceria nanoparticles have
recently been synthesised with {100} at each of the six surfaces.12,13,14
It is worth noting that the predictive capability of simulation has
proven to be even more relevant now than when these calculations
were performed over 15 years ago.15

5.2.5 Surface oxygen vacancies


Equipped with surface models, one can determine those surfaces
that are catalytically most active in that they proffer labile oxygen
species. For the atomistic simulator this translates into the question:
which surface has the lowest oxygen vacancy formation energy?
Such direct interpretation is a view not necessarily held by the
experimentalist as the simulator views the problem with respect to
an atomistically perfect, pristine material.
Calculations using such pristine models have been performed
and revealed that the (110) and (310) surfaces are more reactive
than (111). The result is perhaps intuitive because the coordination
number of ions at the surface of (110) and (310) is lower than
(111). Thus, simulation was able to offer an important prediction:
‘selectively expose the (110) or (310) surfaces to make a more reac-
tive catalyst’. However, one is left with a conundrum because the

b1469_Ch-05.indd 253 4/8/2013 12:31:05 PM


b1469 Catalysis by Ceria and Related Materials

254 D. C. Sayle and T. X. T. Sayle

more reactive surfaces are also less stable and therefore less likely to
be exposed in preference to the more stable (and less reactive)
(111). On the other hand, various methods are available to help
selectively expose the most active surfaces, including supporting the
CeO2 on a substrate, which helps template the evolution of more
reactive ceria surfaces, or, more recently, by working at the
nanoscale.
In an attempt to explore how chemically reactive surfaces might
be exposed in preference to less reactive, but thermodynamically
more stable surfaces, thin films of ceria were supported on an
yttrium-stabilised zirconia (YSZ) (110) surface. The rationale under-
pinning the strategy was that the (110) surface expressed by the YSZ
substrate might template the ceria thin film to also express the (110)
surface in preference to the less reactive (111) surface as both the
ceria and YSZ have a fluorite crystal structure. The simulations
revealed that although the CeO2(110) surface was indeed expressed
at the surface of the thin film, the CeO2(111) and (100) surfaces
were also observed. In particular, the ceria thin film comprised
many misoriented nanocrystalline grains. This was attributed to the
lattice misfit between the ceria and underlying YSZ substrate.
It is perhaps gratifying to note that the ‘perceived’ simplicity of
ceria surfaces, surface steps and surface vacancies, based upon
inspection of the atomistic models, for example Fig. 5.2, are not that
dissimilar to experimental images measured more recently, Fig. 5.3,
using AFM.16,17

5.2.6 Surface oxygen transport


Once oxygen has been extracted from the surface, the material
needs to replenish the supply. Accordingly, one is interested in the
transport of oxygen to and from the surface. A surface is structurally
different compared to the pristine bulk material, which has implica-
tions for the transport of oxygen at the surface or near surface
regions. Indeed, Baudin and co-workers used MD simulation to
explore oxygen mobility at the near surface region of Ca-doped
CeO2(011) and found it to be much faster compared with transport
in the bulk.18

b1469_Ch-05.indd 254 4/8/2013 12:31:05 PM


b1469 Catalysis by Ceria and Related Materials

Atomistic Modelling of Ceria Nanostructures: Introducing Structural Complexity 255

(a) (b)

(c) (d)

(e) (f)

Figure 5.3 (a), (b) Step edges on a CeO2(111) surface as revealed by dynamic scan-
ning force microscopy (SFM). Reproduced with permission from Gritschneder et al.17
(c)–(f) Surface vacancies appearing as (c) single vacancies, (d) in-line configurations
and (e), (f) triangular vacancy clusters, as revealed using high-resolution dynamic
scanning force microscopy operated in non-contact mode. Reproduced with permis-
sion from Gritschneder and Reichling.16 Copyright 2007 IOP Publishing Ltd.

b1469_Ch-05.indd 255 4/8/2013 12:31:05 PM


b1469 Catalysis by Ceria and Related Materials

256 D. C. Sayle and T. X. T. Sayle

In addition, a recent study by Li and co-workers used first-


principle calculations to explore the diffusion of oxygen on a
CeO2(111) surface and identified a two-step surface mechanism,
which is energetically favoured over the most common ‘hopping’
mechanism.19 Indeed, they predicted that the two-step exchange
mechanism would occur ‘eight orders of magnitude faster than that
of the hopping mechanism at 500 K based on the diffusion barriers’.
The MD and first-principles studies revealed that the diffusion
mechanism is critically dependent on the structure at the surface
and therefore such predictions are unlikely to be universally appli-
cable to all ceria surfaces. In particular, some surfaces, for example,
dipolar, will undergo more substantive relaxation (even faceting)
compared to other surfaces and therefore one needs to simulate O
transport for each surface individually before trends can be identi-
fied.20 Clearly, computer simulation is well placed to explore such
diffusion and determine the activation energy barriers associated
with O transport to particular surfaces.

5.2.7 Dislocations
A dislocation (necessarily) terminates at a free surface. Accordingly,
similar to a surface step, the coordination number and local envi-
ronment of surface oxygen in the core region of a dislocation differs
to that of surface oxygen ions far from the core,21 Fig. 5.4, which will
influence the lability of surface oxygen at or near dislocation cores.
Similarly, the local environment of an oxygen ion in the core of the
dislocation far below the surface is also different compared to oxy-
gen residing in (pristine) bulk regions, with implications for ionic
transport in the material. Accordingly, if simulation is to be used to
explore the catalytic activity of oxides, which rests upon surface
lability and ionic transport, it is important to include dislocations
in the atomistic model. The influence of dislocations on point
defects in ionic oxides has been explored by Zhang and co-workers
who constructed atomistic models of dislocations in MgO and
found that defects are strongly bound to the dislocation cores.22
Specifically, vacancies are found to be most stable when they remove

b1469_Ch-05.indd 256 4/8/2013 12:31:05 PM


b1469 Catalysis by Ceria and Related Materials

Atomistic Modelling of Ceria Nanostructures: Introducing Structural Complexity 257

Figure 5.4 Atomistic structure of a screw-edge dislocation in MgO. The atoms


comprising the core, which protrude out of the surface, are shown as dark (oxygen)
and light (magnesium) spheres. Reproduced from Sayle21 with permission from the
Royal Society of Chemistry.

under-coordinated ions at the tip of the extra half plane. They also
found that vacancy migration along the dislocation line will be sub-
stantially enhanced compared to migration through the dislocation-
free crystal structure. On the other hand the mechanisms for oxygen
transport in ceramic oxides are material (crystal structure) specific.
Accordingly, one cannot extend trends identified for one material
to other material classes; rather all one can safely attest is that the
strain fields, emanating from dislocations, will influence surface
oxygen vacancy formation and mobility with important conse-
quences for catalysis. Such calculations can now be explored rou-
tinely using the GULP code.3

5.2.8 Grain boundaries


Inspection of the atomistic structure of a grain boundary (GB)
reveals that ions in the GB region can exist in very different

b1469_Ch-05.indd 257 4/8/2013 12:31:05 PM


b1469 Catalysis by Ceria and Related Materials

258 D. C. Sayle and T. X. T. Sayle

(a) (c) (e) (a) (b)

(c)
(b) (d) (f)

Figure 5.5 Images of a [001](210) Σ5 grain boundary in a CeO2 thin film. Left:
Scanning tunnelling electron micrographs. Right: Atomistic models. Reproduced
with permission from Hojo et al.23 Copyright 2010 American Chemical Society.

environments compared to ions in the pristine bulk material; the


atomistic structure of a GB in ceria is shown in Fig. 5.5. Accordingly,
similar to dislocations, grain boundaries can influence the energet-
ics and transport properties of oxygen in ceria. There are several
atomistic simulation codes that use symmetry operators to construct
atomistic models for grain boundaries, see for example METADISE.9
The easiest to construct are high symmetry boundaries, such as
twins, because the unit cells are necessarily smaller compared to
more general grain boundaries. In particular, an atomistic model of
a (210) Σ5 grain boundary in ceria has been modelled, Fig. 5.5, with
excellent accord to experiment.23 Moreover, the study revealed that
oxygen vacancies were found to have a pivotal role in the stability of
the grain boundary. Indeed, Tasker showed that the energy of a
grain boundary can be lowered with the inclusion of vacancies.
Accordingly he argued that the vacancies are not defects as such;
rather ‘an intrinsic part of the low-temperature structure’.24

b1469_Ch-05.indd 258 4/8/2013 12:31:05 PM


b1469 Catalysis by Ceria and Related Materials

Atomistic Modelling of Ceria Nanostructures: Introducing Structural Complexity 259

Oxygen mobility in grain-boundary regions can be calculated in


an analogous fashion to O mobility in the bulk crystal, for example,
see Duffy and Tasker.25 In particular, Fisher and Matsubara gener-
ated atomistic models of Σ5 (310)/[001], Σ13(320)/[001] and Σ5
(111) 60ο grain boundaries in 8 mol% YSZ and found that oxygen
diffusion in the systems including grain boundaries showed that the
tilt grain boundaries reduce the overall ionic conductivity relative to
a single crystal of 8 mol% YSZ.26 Conversely, the Σ5 twist boundary is
able to support rapid diffusion and increases total conductivity. The
results are important in that the work shows that generalised state-
ments such as ‘grain boundaries either enhance or inhibit oxygen
mobility’ are not realistic; rather the atomistic structure of the grain
boundary is pivotal to oxygen mobility along it and therefore must
be considered on a case-by-case basis; a concise account of model-
ling and simulation of grain boundaries is given in the advanced text
by Sutton and Balluffi.27
Similar to grain boundaries, the influence of dislocations, as
argued above, will necessarily need to be considered individually.
One might then also question the influence of a dislocation and
neighbouring grain boundary. Could the influence of both reverse
trends by either the dislocation or grain boundary in isolation? The
complexity of such defects suggests that this is not an unreasonable
question. Accordingly, if one is to calculate the influence of such
defects, then one needs to include both in a single structural model
together with the synergy of interaction.

5.2.9 Supported thin films and interfaces


Supporting a material on a host substrate can lead to a change in the
structure and hence (catalytic) properties of the supported mate-
rial. In particular, there normally exists a misfit between the lattice
parameter of the thin film and the underlying substrate. Accordingly,
a supported thin film can be either stretched or compressed to
facilitate a good ‘epitaxial’ match across the interface. Clearly, com-
pressing or tensioning a thin film will affect the energy required to
extract oxygen from the surface of the thin film and therefore

b1469_Ch-05.indd 259 4/8/2013 12:31:06 PM


b1469 Catalysis by Ceria and Related Materials

260 D. C. Sayle and T. X. T. Sayle

consideration of thin films is of importance in the field of catalysis.


Ultimately, one should be able to optimise the (catalytic) reactivity
of a supported thin film by appropriate choice of substrate (misfit)
and thin-film thickness.
Supporting a thin film on a substrate material may also result
in the supported thin film expressing different faces at the surface
compared to the unsupported material, since a supported thin
film attempts to optimise favourable interactions across the inter-
facial region. Such selective exposure can be beneficial in that the
material may express a catalytically more reactive surface.
Accordingly, understanding the science and mechanisms under-
pinning such phenomena is important in the intelligent design of
new catalysts.
Atomistic models for supported thin ceria films, constructed
using a near coincidence site lattice theory,28 revealed that by sup-
porting ceria on an alumina substrate, the oxygen vacancy forma-
tion energy can be reduced compared to the parent (unsupported)
CeO2. This study therefore suggests that it is possible to increase
the reactivity of ceria by supporting a thin film on a substrate. On
the other hand this study is naïvely simplistic in that many other
factors must be considered. This includes, for example, the expo-
sure of particular (reactive) faces — simulation studies have
indicated that supported thin films can influence the surface of
ceria preferentially exposed29 — and the presence of surface steps,
edges, intrinsic and extrinsic defects, dislocations and grain
boundaries.

5.3 Simulating ‘The Dirt’


One of the most powerful attributes of atomistic simulation is its
ability to deconvolute microstructural features and explore the
effect of a particular structural feature, such as a point defect or
dislocation, on the properties in isolation and therefore without
any influence from other (neighbouring) structural features. In
particular, it is far easier to construct a model with just one particu-
lar defect, for example a vacancy, than with multiple defects, such

b1469_Ch-05.indd 260 4/8/2013 12:31:06 PM


b1469 Catalysis by Ceria and Related Materials

Atomistic Modelling of Ceria Nanostructures: Introducing Structural Complexity 261

as vacancies, dislocations and grain boundaries, within a single


atomistic model. On the other hand, the greatest attribute of atom-
istic simulation is also one of its weaknesses. In particular, real
materials comprise a variety of different microstructural features,
which will interact synergisticallya — such synergy can influence,
sometimes profoundly, their properties. To simulate properties
that can be compared directly with experiment, an atomistic
model needs to include all microstructural features together with
their synergy of interaction observed experimentally. Thus the
phrase ‘simulate the dirt’ was coined, which represents the chal-
lenge to the simulator in capturing the structural complexity in a
real system. Certainly, ‘the dirt’ is a wholly vague term and perhaps
inappropriate for scientific description, but it does have value in
that it captures the essence that neither experiment nor simulation
truly understands the complexity with atomistic clarity, but that
such complexity, coined ‘the dirt’, can have a large effect on the
properties.

5.3.1 Ceria nanotubes


The simulations and results described thus far include generally
only one or two structural features within a pristine model. To cap-
ture a rich microstructure in a single model is challenging using
atomistic simulation. However, a model for a ceria nanotube was
generated by Martin and co-workers.30 In particular, this model,
shown in Fig. 5.6, includes the morphology of the tube including
surfaces exposed, steps and edges together with a variety of grain
boundaries and dislocations. Using such a model, the authors were
able to compute the oxygen vacancy formation and found that
vacancies are stabilised at the surfaces, grain boundaries and triple
junctions. The authors also commented that ‘the structure repre-
sents a new nanostructure that, combined with its inherent high

a
For example, the stress field of a dislocation will likely influence the structural
configuration of intrinsic and extrinsic defects localised in or near the core region
of the dislocation.

b1469_Ch-05.indd 261 4/8/2013 12:31:06 PM


b1469 Catalysis by Ceria and Related Materials

262 D. C. Sayle and T. X. T. Sayle

(a) (b) (c)

(d) (e) (f)

Figure 5.6 Top: Atomistic model of a CeO2 nanotube, which includes microstruc-
tural features such as dislocations and grain boundaries. Bottom: The method used
to generate such models. Reprinted with permission from Martin et al.30 Copyright
2007 American Chemical Society.

b1469_Ch-05.indd 262 4/8/2013 12:31:06 PM


b1469 Catalysis by Ceria and Related Materials

Atomistic Modelling of Ceria Nanostructures: Introducing Structural Complexity 263

surface area to bulk ratio, should have enhanced activity, which


would unlock the oxygen storage capabilities of ceria’. This predic-
tion was validated by Zhou and co-workers, who synthesised ceria
nanotubes and found that they ‘exhibit excellent reducibility and
high oxygen storage capacity, indicating they are potential novel
catalytic materials’.31

5.3.2 Outlook
Atomistic simulation methodologies that first construct a crystal of
the pristine material, derive appropriate forcefields and then
introduce microstructural features are well placed to help experi-
ments elucidate phenomena of critical importance for catalysis.
However, the competence of the simulator is necessarily at the
expert level, because they need to be conversant with the workings
of the code and therefore be able to perform and interpret the
data emanating from such methods. To help address such techni-
cal issues, commercial packages are available that enable one to
perform all these simulations at a single ‘point of contact’. For
example, via a graphical interface, one can build the structural
model with template examples, link the structure with appropriate
forcefields, introduce pertinent microstructural features (point
defects, morphology, dislocations, grain boundaries) and then cal-
culate properties (such as oxygen vacancy formation energy) and
simulate pertinent processes (such as oxygen diffusion), using, for
example, energy minimisation or molecular dynamical simulation.
Moreover, this can all be achieved using appropriate ‘clicks of a
mouse button’. Such ‘simulation studios’ reduce the amount of
time required to extract data, offer high throughput and do not
necessarily require the user to be an ‘expert’. Conversely, limited
knowledge of the workings, subtleties and science underpinning
such studios can potentially lead to naïve or indeed incorrect
information being computed. Accordingly, one must advocate care
in the utilisation of such methods and be critically aware of the
pitfalls.

b1469_Ch-05.indd 263 4/8/2013 12:31:07 PM


b1469 Catalysis by Ceria and Related Materials

264 D. C. Sayle and T. X. T. Sayle

5.4 Nanocatalysis
A recent major scientific finding reveals that one can modify or
indeed control the physical, chemical and mechanical properties of
a material simply by limiting the size of the material to the nanome-
tre range. Indeed, it has been proposed that nanomaterials will
transform the scientific landscape of the 21st century. Certainly, its
influence on catalysis cannot be understated as evidenced by an
exponential growth in publications in this area, proffering new
physical phenomena, with exploitation of the science giving rise to
new and proven applications. If atomistic simulation is to continue
to provide valuable insights and predictions, one must strive to gen-
erate models for nanomaterials.
One of the benefits of the ‘nanoscale’ is that atomistic simula-
tion can simulate every atom comprising the nanomaterial explic-
itly, without resorting to symmetry methods to facilitate periodic
boundaries. On the other hand, it brings new hierarchical levels of
structural complexity. For example, to simulate a nanomaterial, one
must capture in the atomistic model:

• the crystal structure,


• the microstructure including: defects, dislocations and grain
boundaries,
• the nanostructure that is the morphology including the surfaces
exposed, steps, edges and vertices together with issues of curva-
ture (convex and concave).

Perhaps more importantly, each structural level will be synergis-


tically related and therefore all structure features must be included
in a single structural model to capture the synergy and the influence
this has over the entire atomistic model.
It is difficult to construct atomistic models for a nanomaterial
using conventional methods of simulation, such as symmetry opera-
tors, because of the hierarchical levels of structural complexity.
Experimentally, such nanostructures evolved during synthesis and
therefore perhaps the easiest way to generate full atomistic models
is to ‘simulate synthesis’. We explore such endeavours in the second
section of this chapter.

b1469_Ch-05.indd 264 4/8/2013 12:31:07 PM


b1469 Catalysis by Ceria and Related Materials

Atomistic Modelling of Ceria Nanostructures: Introducing Structural Complexity 265

5.5 Simulating Synthesis


The essential difference between experiment (synthesis) and atom-
istic simulation is the ability of the simulator to be able to manipu-
late the structure with total atomistic control. Conversely, the
experimentalist has less control and is generally limited to being
able to manipulate only the reagents and synthetic conditions of
manufacture. As shown later, attempts to ‘simulate synthesis’ to gen-
erate atomistic models with sufficient structural detail and complex-
ity to be of value to experimenters have been successful. However,
this does come at a cost in that ‘atomistic modelling and simulation’
looses, in part, potentially one of its most important attributes —
total atomistic control.
In this second section of this chapter we explore simulation
methods, which can loosely be described as ‘simulating synthesis’
including: simulated annealing, atom deposition and simulated
crystallisation.

5.5.1 Simulated annealing and Monte Carlo


Nanoparticles of ceria comprising up to 150 atoms were generated
by Cordatos and co-workers using simulated annealing comple-
mented by a standard metropolis walk.32 In particular, the Ce and O
atoms comprising the nanoparticle were introduced randomly into
a simulation cell and each atom moved at random in order to drive
the system to a lower energy configuration. Moves were either
accepted or rejected following a Monte Carlo approach, which
helped the ions to overcome high energy barriers associated with
local minima or metastable configurations and enabled global
energy minima configurations to be reached. The structure for a
150-atom nanoparticle, which evolved during the annealing process,
Fig. 5.7, exposed a (111) surface facet, indicating that this is the
most stable surface. Moreover, the calculations revealed that the
energy required to reduce the cluster generally increased with
increasing cluster size. The important prediction emanating
from this study, is that smaller (nanoparticles) are catalytically more
reactive with respect to oxidation reactions.

b1469_Ch-05.indd 265 4/8/2013 12:31:07 PM


b1469 Catalysis by Ceria and Related Materials

266 D. C. Sayle and T. X. T. Sayle

Figure 5.7 CeO2 cluster comprising 50 formula units; the (111) plane is high-
lighted. Reprinted with permission from Cordatos et al.32 Copyright 1996 American
Chemical Society.

Pivotal to this study was that the nanoparticle structure was not
influenced (perhaps erroneously) by the simulator constructing
configurations that appear intuitively feasible according to, for
example, the crystal structure or likely surfaces exposed; rather the
configurations were driven energetically. Moreover, it is gratifying
to see that the lowest energy surface of ceria, CeO2(111), indeed
evolved using this method.

5.5.2 Atom deposition


Experimentally one can fabricate advanced materials by depositing
a material on a host lattice. In particular, ceria thin films, supported
on a variety of substrates, have been synthesised using, for example,
molecular beam epitaxy and chemical vapour deposition. Here, the

b1469_Ch-05.indd 266 4/8/2013 12:31:07 PM


b1469 Catalysis by Ceria and Related Materials

Atomistic Modelling of Ceria Nanostructures: Introducing Structural Complexity 267

influence of the substrate can be used to introduce strain and struc-


tural modification into the thin film owing to the lattice misfit
between the two materials and the interactions across the interface.
Such strain-induced structural transformations are well known to
influence the catalytic properties profoundly.
Atomistic simulation can also be used to simulate this process.
One particular technique involves firing ions from a source onto the
surface of the substrate, Fig. 5.8. This strategy has been used to gen-
erate atomistic models of thin oxide films on oxide substrates. In
particular, the simulations revealed the embryonic stages of nuclea-
tion and crystal growth of thin films.33 The lattice misfit was also
shown to have a profound influence over the structure leading to
the evolution of strain within the thin film, which was eventually
quenched via the evolution of dislocations at particular ‘critical’
thicknesses. The idea of critical areas (for dislocation evolution) was
also proposed.34 Clearly, the evolution of microstructural features
and localised strain arising to help accommodate the lattice misfit is
important to the catalytic properties of the material and therefore
insight into the behaviour at the atomistic level, as derived from
simulation, can help identify improved materials.
The effect of supporting a thin film on a substrate can also lead
to a variety of (potentially catalytically reactive) surfaces being
exposed by the thin film.b In particular, Fig. 5.8(c) shows a heavily
corrugated CaO surface, which at a particular ‘critical’ thickness
transforms via dislocation evolution to CaO(100), Fig. 5.8(d).
Accordingly, this paves the way for ‘computer aided materials
design’ in that various substrates and thin-film thicknesses can be
explored computationally to determine systems with desirable (reac-
tive) surfaces; such predictions can then be tested experimentally.
A third approach is to simulate crystallisation. In particular, all
(crystalline) material synthesis involves some kind of crystallisation
process. This may be crystallisation in solution or crystallisation

b
The energetically most stable ceria surface CeO2(111) is catalytically less reactive
compared to CeO2(110) or CeO2(100), yet is preferentially exposed in the parent
bulk material because of its inherent stability.

b1469_Ch-05.indd 267 4/8/2013 12:31:07 PM


b1469 Catalysis by Ceria and Related Materials

268 D. C. Sayle and T. X. T. Sayle

(a) (b)

(c) (d)

Figure 5.8 Atomistic models of supported thin films generated by simulating atom
deposition onto a substrate. (a) Illustration of the process. (b) The mobility of the
ions once deposited onto the surface. (c) Atomistic model of a CaO thin film sup-
ported on MgO(100) after the deposition of 1.8 equivalent monolayers onto the
surface. (d) After 3.6 equivalent monolayers have been deposited onto the surface.
Note in (d) the presence of a mixed screw-edge dislocation (arrow), which evolves
in the supported CaO thin film to accommodate the lattice misfit. Reproduced
from Sayle et al.33 with permission from the Royal Society of Chemistry.

from a molten precursor (solidification). We will consider arguably


the easiest — crystallisation from liquid to solid, although a beauti-
fully elegant example of crystallisation of ions in solution is given in
Piana et al.35 and there are currently considerable interests and
efforts directed in this area.

b1469_Ch-05.indd 268 4/8/2013 12:31:07 PM


b1469 Catalysis by Ceria and Related Materials

Atomistic Modelling of Ceria Nanostructures: Introducing Structural Complexity 269

5.5.3 Simulated crystallisation


Crystallisation (solidification) is a phase change from a liquid to solid
state and involves the atoms rearranging periodically with long-range
order. However, crucially during crystallisation, microstructural fea-
tures can evolve within the solidifying material including dislocations,
grain boundaries, defects and surfaces exposed (morphology).
Accordingly, if one could simulate crystallisation, then potentially one
might also capture important microstructural features within the
model — together with their synergy of interaction — without having
to introduce the structural features ‘artificially’ using, for example,
symmetry operators. Specifically, if one were able to simulate a process,
which occurs during synthesis to produce the complex microstructure,
then arguably the microstructural features should be an accurate
reflection (structurally) of the real material because the models were
generated by simulating their experimental mode of manufacture.
A major problem associated with simulating crystallisation
directly using MD is that the energy barriers associated with evolving
a nucleating seed may be high. Moreover, the number of possible
non-crystalline (amorphous) configurations is high, which necessi-
tates sampling a large number of configurations before reaching a
‘crystalline’ configuration. Accordingly, the probability of sampling
trajectories that lead to a crystalline structure is low considering the
small windows of time accessible to MD, which typically are of the
order of nanoseconds. In practice, this can lead to many hours of
‘wasted’ computational effort simply exploring a configurational
space without ever evolving a nucleating seed; the structure remains
amorphous at the end of the simulation. Similar to experiments,
one can modify the simulation conditions such as temperature or
pressure. For example, increased temperature can help the atoms
overcome high energy barriers and thus speed up the simulation;
similarly, increases in pressure favour more dense (crystalline) struc-
tures. On the other hand high temperatures may lead to facile rea-
morphisation of crystalline seeds if they did evolve. There are several
approaches that have been devised in order to circumvent such
challenges, two of which are described below:

b1469_Ch-05.indd 269 4/8/2013 12:31:09 PM


b1469 Catalysis by Ceria and Related Materials

270 D. C. Sayle and T. X. T. Sayle

5.5.3.1 Seeded crystallisation


Seeded crystallisation entails introducing a ‘pre-crystallised’ seed
(perhaps generated using symmetry) into an amorphous sea of
ions.36 The seed then nucleates the crystallisation of the amorphous
regions.
A strategy, which is perhaps more aligned with experiment, is
simulated templated crystallisation. Here, one uses a (crystalline)
substrate material to template the crystallisation of an overlying
amorphous thin film.37 In particular, Fig. 5.9 shows the structure of
a thin film of MgO supported on BaO.38 Such simulations, similar to
atom deposition, attempt to simulate processes that occur during,
for example, chemical vapour deposition or molecular beam epi-
taxy, where at some point the substrate will help nucleate and
template the crystallisation of the amorphous thin film deposited
thereon.

5.5.3.2 Bias molecular dynamics


Another approach to help facilitate crystallisation is to bias the
trajectories of all the atoms moving in the (molten/amorphous)
configuration to favour crystallisation. Specifically, order parameters
(which can be bond distances, coordination numbers, nearest
neighbour densities and radial distribution functions) may be used
as a gauge of crystallinity and used to help drive the simulation (tra-
jectory) to favour maximising the order parameter and, ultimately,
induce crystallinity.
One of the problems associated with these types of simulations,
is that MD trajectories, which do not lead to crystalline structures,
may be explored repeatedly. To prevent such occurrence and
reduce computational effort, a trajectory history can be recorded
and compared with the current trajectory. If a match is found, then
the trajectory can be modified to ensure that the simulation always
explores a new phase space and structure. Clearly this approach can
be overwhelmed by the sheer number of configurations that are pos-
sible. However, the technique has been used successfully by Quigley

b1469_Ch-05.indd 270 4/8/2013 12:31:09 PM


b1469 Catalysis by Ceria and Related Materials

Atomistic Modelling of Ceria Nanostructures: Introducing Structural Complexity 271

(a) (b)

(c) (d)

Figure 5.9 MgO thin film supported on a BaO(001) substrate. (a) Plan view look-
ing down onto the MgO thin film showing the polycrystallinity of the material.
(b) Segment of (a) showing more clearly the structure between two misoriented
MgO grains. (c) Enlarged segment of (b) showing the dislocation core.
(d) Perspective view of the MgO/BaO(001) system — only two layers of the BaO
substrate are shown to improve clarity. Reproduced from Sayle and Watson39 with
permission from the Royal Society of Chemistry.

and Rodger to model the crystallisation of water into ice39 as shown


in Fig. 5.10.
There are a wealth of other innovative techniques to circumvent
such challenges including accelerated dynamics, metadynamics and
hyperdynamics. A review of the challenges and state of the art of

b1469_Ch-05.indd 271 4/8/2013 12:31:09 PM


b1469 Catalysis by Ceria and Related Materials

272 D. C. Sayle and T. X. T. Sayle

Figure 5.10 Molecular dynamics simulation of water freezing at 180 K showing the
growth of a critical nucleus as a function of simulation time. Hydrogen bonds
connecting molecules, identified as solid, are highlighted. Reproduced from
Quigley and Rodger40 with permission from the Taylor & Francis Group.

b1469_Ch-05.indd 272 4/8/2013 12:31:12 PM


b1469 Catalysis by Ceria and Related Materials

Atomistic Modelling of Ceria Nanostructures: Introducing Structural Complexity 273

crystallisation simulation at the atomistic level is given by Anwar and


Zahn40 although a comment from this review was that ‘one does not
get something for nothing’. Specifically, any form of bias introduced
to accelerate or direct the dynamics may also introduce a bias in the
final crystalline structures and therefore render the structures
derived from the method, in part, artificial.
A milestone in the simulation of crystallisation was performed by
Streitz and co-workers, who simulated the crystallisation (solidifica-
tion) of a metal without size effects.41 In particular, they simulated
the spontaneous nucleation and growth of a solid from the liquid
phase. The authors found that 16 million atoms were sufficient to
simulate metal solidification from the melt with no approximations
due to finite system size. Equipped with an IBM BlueGene/L com-
puter at LLNL they were able to simulate a billion atoms. On the
other hand those of us without such resources can still simulate
nanoparticles as one can easily consider all the atoms comprising
the nanoparticle explicitly.

5.5.4 Ceria nanoparticles and nanorods


To generate models for ceria nanoparticles, Sayle and co-workers
melted ca. 16,000 atoms of ceria by applying molecular dynamics
simulation at high temperatures (7000 °C). The system was then
cooled (under MD) to a temperature below its melting point. At a
particular point in time, a crystalline seed (with fluorite structure)
spontaneously evolved within the molten sea of ions and nucleated
crystallisation of the surrounding (molten) atoms,42 Fig. 5.11.
If the ceria nanoparticle was doped with Ti, then the final nano-
particle shape was spherical43 compared to the octahedral morphol-
ogy of the un-doped CeO2. This was because the TiO2 surrounded
the nanoparticle during crystallisation and remained molten at
the simulation temperature (and hence imposed sphericity on the
nanoparticle to minimise the surface energy). Accordingly, the
ceria necessarily crystallised from the inside out. Conversely for
the undoped material, the ceria was able to crystallise from the

b1469_Ch-05.indd 273 4/8/2013 12:31:12 PM


b1469 Catalysis by Ceria and Related Materials

274 D. C. Sayle and T. X. T. Sayle

(a) (b) (c) (d)

(e) (f) (g)

(h) (i) (j)

(k) (l)

Figure 5.11 Atomistic structures of ceria nanoparticles. (a)–(d) Snapshots taken


during an MD simulation showing the crystallisation of a Ti–CeO2 nanoparticle
from a molten precursor. (e) Model atomistic structure of a CeO2 nanoparticle as
it starts to solidify; the oval indicates the nucleating seed, which spontaneously
evolves on the surface. (f) Surface rendered model of a CeO2 nanoparticle after

b1469_Ch-05.indd 274 4/8/2013 12:31:12 PM


b1469 Catalysis by Ceria and Related Materials

Atomistic Modelling of Ceria Nanostructures: Introducing Structural Complexity 275

outside in by evolving a nucleating seed that comprised and


exposed the stable (111) surface leading to a polyhedral morphol-
ogy with {111} facets truncated by {100} in quantitative agreement
with experiment.43 This strategy also enables one to compare the
atomistic model with the real material directly, Fig. 5.11. We also
note that similar to the real materials, the model nanoparticles
comprised a rich microstructure including morphology and sur-
faces exposed, point defects, dislocations and grain boundaries
(Fig. 5.11).
A similar procedure was used to generate ceria nanorods.46,47 In
particular, amorphous nanoparticles were allowed to agglomerate in
one direction resulting in amorphous nanorods, which were then
crystallised. Specifically, nanorods with growth directions along
[110] and [211], Fig. 5.12, were evolved by performing the crystal-
lisation step at different temperatures. For the nanorod with the
[110] growth direction, {111} surfaces were predominantly exposed.
Conversely, for the [211] growth direction, both {111} and {100} sur-
faces were exposed; the latter were also facetted to {111} to help
quench the dipole associated with {100}, which led to the corrugated
structure of the nanorod, Fig. 5.12. A third nanorod evolved two
nucleating seeds during crystallisation. Accordingly, as these (neces-
sarily) misoriented seeds nucleated the crystallisation of the sur-
rounding amorphous sea of ions, they impinged on each other

Figure 5.11 (Continued) crystallisation. (g) 3D tomogram of a CeO2 nanocrystal


generated from computer-aided tomography of sequential oriented transmission
electron microscopy (TEM) images. (h) Model atomistic structure of a Ti–CeO2
nanocrystal showing the embryonic nucleating seed, which evolves inside the molten
nanoparticle. (i) Surface rendered model of a (core-shell) Ti–CeO2 nanoparticle
after crystallisation. (j) high resolution TEM (HRTEM) image of a Ti–CeO2 nanocrys-
tal revealing the (amorphous) TiO2 shell, which encapsulates the inner (crystalline)
CeO2 core rendering it spherical. (k) Slice through a (model) CeO2 nanoparticle
which comprises various misoriented CeO2 grains. (l) Enlarged segment of (k) reveal-
ing the atomistic structure of a grain boundary. (a)–(f), (h)–(j) Science 312, 1504,
2006. Reprinted with permission from AAAS from Feng et al.43 (g) Reprinted with
permission from Sayle et al.44 Copyright 2008 American Chemical Society. (k),(l)
Reproduced from Sayle et al.45 with permission from the Royal Society of Chemistry.

b1469_Ch-05.indd 275 4/8/2013 12:31:32 PM


b1469 Catalysis by Ceria and Related Materials

276 D. C. Sayle and T. X. T. Sayle

(a) (b)

(c) (d) (e)

Figure 5.12 Atomistic models and HRTEM images of ceria nanorods. (a) and (b)
Model CeO2 nanorods, which extend along [110] and (b), respectively. (c) Nanorod
with twin-grain boundaries; the atomistic structure of the grain-boundary region is
shown enlarged in the inset. HRTEM images of nanorods with [110] and [211] growth
directions are shown in (d) and (e), respectively. (a)–(c) Reprinted with permission
from Sayle et al.46 Copyright 2007 American Chemical Society. (d), (e) Reprinted with
permission from Du et al.48 Copyright 2007 American Chemical Society.

resulting in the formation of a twin boundary, Fig. 5.12(c).


Experimentally, nanorods with [110] and [211] growth directions
have been synthesised; HRTEM are shown in direct comparison to
the atomistic models.48
It can be argued that these models are easier to construct, com-
pared to strategies that exploit symmetry operators as described in the
first half of this chapter. In particular, one needs only to specify a (crys-
tallisation) temperature and molecular dynamical simulation directs
the atoms into appropriate low-energy configurations, which include
the evolution of a rich microstructure. However, it can be difficult to
ascertain the (simulation) conditions that facilitate crystallisation.
A comprehensive review on modelling and simulation of nano-
particles is given by Barnard49 and Catlow and co-workers.50

b1469_Ch-05.indd 276 4/8/2013 12:31:32 PM


b1469 Catalysis by Ceria and Related Materials

Atomistic Modelling of Ceria Nanostructures: Introducing Structural Complexity 277

5.5.4.1 Oxidation of CO to CO2 using ceria nanoparticles and nanorods


Equipped with such models one can begin to answer whether these
nanomaterials are catalytically more active compared with the
parent bulk material. Simple inspection can give some clues. In par-
ticular, the surface of the ceria is pivotal to its reactivity — it was
shown that CeO2(310) (step surface) is more reactive than CeO2(111)
as the oxygen ions on the surface of a step have necessarily less coor-
dination compared with oxygen on CeO2(111). However, CeO2(111)
is more stable and therefore preferentially exposed; for the parent
bulk material the challenge is to facilitate exposure of chemically
more reactive surfaces. For the nanoparticle, its small size means
that the concentration of steps, and hence labile oxygen species, is
increased compared to the parent bulk material. Accordingly, tra-
versing to the nanoscale affords a method of creating more catalyti-
cally active surfaces. Moreover, calculations revealed that the surface
oxygen is more labile compared to oxygen species on the equivalent
surface of the parent bulk material.51 In particular, a ceria nanopar-
ticle was used to calculate the energy required to convert CO to CO2
using oxygen taken from the surface and subsurface of the ceria
nanoparticle. Specifically:

CO(g) + CeO2(nanoparticle) Æ CO2(g) + Voii + 2Ce¢Ce

The reactivity of ceria nanoparticles has also been computed


using quantum mechanical methods. In particular, Migani and co-
workers considered (CeO2)n nanoparticles with sizes n = 21, 30, 40
and 80. Their calculations revealed that oxygen extraction from
the surface of the nanoparticle is easier than oxygen extraction
from the extended (‘even irregular’) surface of the parent bulk
material.52 Moreover, the authors found that an increase of particle
size is accompanied by a dramatic decrease of oxygen formation
energy, implying that at certain sizes this energy should reach a
minimum.
Similarly, one can consider the reactivity of ceria nanorods. Here,
the simulations revealed that the growth directions of ceria nanorods
are important with respect to reactivity as rods with different growth

b1469_Ch-05.indd 277 4/8/2013 12:31:39 PM


b1469 Catalysis by Ceria and Related Materials

278 D. C. Sayle and T. X. T. Sayle

directions necessarily expose different proportions of the reactive


surfaces. In particular, the {100} surfaces exposed by the nanorod,
Fig. 5.12, are likely more reactive owing to the comparative ease of
extracting oxygen from this surface compared to {111} surfaces.
Simulation can therefore provide valuable insight for experiment.
For example, if an experiment is able to selectively synthesise nanorods
with [211] growth directions, simulations predict that they will likely
be more reactive compared to nanorods aligned along [110], which
expose a lower (area) concentration of {100}.
These findings are supported experimentally.53 In particular, Liu
and co-workers found that ceria nanorods display high oxidative
capability, which may be explained by the high exposure of reactive
surfaces, steps edges, etc. In addition, a study on ceria nanocubes
exposing six CeO2{100} surfaces revealed enhanced oxygen stor-
age,14 which may be attributed to the ease of O extraction from
these surfaces.

5.5.4.2 Oxygen transport in nanoceria


An important process for ceria in catalysis is the mobility of oxygen
within the lattice. Clearly, if oxygen at the surface is used in a
chemical reaction, vacancies will be left on the surface. Indeed, one
of the remarkable properties of ceria is its ability to withstand high
oxygen depletion while retaining the fluorite crystal structure.
Accordingly, to continue the catalytic cycle, the vacancies must be
filled — such cycling is critically dependent upon oxygen conductiv-
ity in the lattice.
A seminal paper by Sata and co-workers showed that ion conduc-
tivity in BaF2/CaF2 heterolayers is tuneable by changing the thick-
ness of the individual BaF2 or CaF2 films; optimum conductivity was
achieved for films with nanometre thicknesses.54 Computer simula-
tions on this system revealed that the F ions diffuse preferentially at
interfacial and grain-boundary regions, which helped explain the
observed thickness-tuned conductivity.55
Although oxygen transport in ceria is vacancy driven, tuneable
oxygen conductivity is tantalisingly attractive — having implications

b1469_Ch-05.indd 278 4/8/2013 12:31:39 PM


b1469 Catalysis by Ceria and Related Materials

Atomistic Modelling of Ceria Nanostructures: Introducing Structural Complexity 279

not only for catalysis but for many other applications of ceria (nota-
bly in solid oxide fuel cells). Accordingly, ceria heterolayers, inter-
faced with YSZ thin films, were constructed using simulated
crystallisation and the ionic conductivity was calculated. During
crystallisation, the ceria evolved misfit dislocations to help accom-
modate the strain in the system associated with the lattice misfit
between the ceria and YSZ, Fig. 5.13. In this instance, the disloca-
tions evolved during the crystallisation and were not introduced
‘manually’ via symmetry operators based upon the relative lattice
parameters.
However, the introduction of such microstructural complexity
led to a significant reduction in the conductivity of oxygen. In par-
ticular, the simulations revealed that O conduction in the ceria was

CeO2

YSZ

Figure 5.13 Atomistic model of a CeO2 thin film supported on YSZ. Top:
Perspective view showing the atomistic structure; a dislocation is shown in the white
oval. Bottom: Side view revealing the periodicity of the misfit dislocations.
Reproduced from Sayle et al.56 by permission of the Royal Society of Chemistry.

b1469_Ch-05.indd 279 4/8/2013 12:31:39 PM


b1469 Catalysis by Ceria and Related Materials

280 D. C. Sayle and T. X. T. Sayle

predominantly vacancy driven, as for the parent bulk material, and


diffusion in grain-boundary or interfacial regions was inhibited
rather than promoted56 thus reducing the overall conductivity.
On the other hand it may be questioned whether the strain in the
thin films (to accommodate the lattice misfit between ceria and
YSZ) is optimised and whether alternative (substrate) materials
need to be used, in conjunction with control over the thin-film
thickness, to facilitate a tuneable strain in the ceria, which promotes
O conduction.
Figure 5.14 shows the relative conductivity of oxygen in various
ceria nanostructures and shows that the fastest conductivity is attrib-
uted to the bulk material.45 Such results cast some doubt as to
whether traversing to the nanoscale will proffer increased conductiv-
ity in ceria. Conversely, a related study by Sankaranarayanan and

0.4 0.5 0.6 0.7 0.8


-14

CeO1.97 parent (bulk)


-15

CeO1.97 thin film

-16
Ln(Di)

Pure CeO2
-17 nanoparticle

CeO1.95
nanoparticle
-18
Rh0.1Ce0.9O1.95
nanoparticle

-19
1000/T

Figure 5.14 Oxygen diffusion coefficients (Di, cm2s−1), calculated as a function of


temperature. Reproduced from Sayle et al.45 by permission of the Royal Society of
Chemistry.

b1469_Ch-05.indd 280 4/8/2013 12:31:40 PM


b1469 Catalysis by Ceria and Related Materials

Atomistic Modelling of Ceria Nanostructures: Introducing Structural Complexity 281

Ramanathan76 on MgO-supported YSZ thin films, revealed a two


orders of magnitude increase in the ionic conductivity in YSZ as the
YSZ film size decreased from 9 to 3 nm owing to a decrease in the
activation energy barrier from 0.54 to 0.35 eV in the 1200–2000 K
temperature range. Simulated amorphisation and crystallisation
were used to generate the atomistic model, which enabled the
authors to capture the microstructure — particularly a grain-bound-
ary network.
Whether the strain induced by the presence of interfaces (rather
than dopants) can be exploited to increase ionic conductivity is still
(at the time of writing) hotly debated.

5.5.5 Mesoporous ceria


Mesoporous ceria with crystalline cell walls57 offers the possibility of
tuneable reactivity as a function of architecture. In particular, one
has the potential to structurally engineer (internal) surfaces with
concave curvature in addition to the necessarily convex curvature
associated with nanoparticles and nanorods. Experimentally, eluci-
dation of the internal surfaces of a mesoporous oxide is challenging;
3D tomography yields remarkable insights.58 Conversely, equipped
with a full atomistic model, molecular graphics visualisation can be
used to explore the internal architecture and surfaces expressed.
However, the generation of such a model is difficult because one has
to consider constructing the architecture of the mesoporous mate-
rial in addition to the crystal structure and microstructure — similar
to generating atomistic models of nanoparticles, perhaps the easiest
way is to copy experiments. To this end one needs to explore syn-
thetic methods of fabrication and ascertain which part(s) of the
synthetic process are most amenable to simulate.
Mesoporous oxides have been fabricated using a variety of meth-
ods spanning templated self-assembly of nanoparticles,59 nanocast-
ing60 and using surfactants to act as a molecular scaffold to help
build the mesoporous architecture.61 Template-free self-assembly of
nanoparticles via ice crystallisation has also been achieved,62 where
the phase change from water to ice ‘sculpts’ the nanoparticles into

b1469_Ch-05.indd 281 4/8/2013 12:31:40 PM


b1469 Catalysis by Ceria and Related Materials

282 D. C. Sayle and T. X. T. Sayle

(a) (b) (c)

(d) (e) (f)

(g) (h)

Figure 5.15 Model atomistic structures and electron tomograms of ceria nano-
chains comprising individual CeO2 nanoparticles. (a), (b) Iso-surface rendered
STEM images of the reconstructed 3D volume of CeO2 nanoparticles comprising a
ceria nanochain. (c) Schematic illustrating the morphology of the individual nano-
particles. (d) HRTEM image of a ceria nanochain revealing the oriented attach-
ment of individual ceria nanoparticles at (100) and (111) faces. (e) Atomistic
model of ceria nanoparticles that (self-) assembled into a nanochain. (f) Iso-surface
rendered model of a ceria nanochain showing oriented attachment at (100) and
(111) faces. (g), (h) Model atomistic structures of individual ceria nanoparticles.
(a), (b), (c) Reprinted with permission from Tan et al.64 Copyright 2011 American
Chemical Society. (d) Reprinted with permission from Du et al.48 Copyright 2007
American Chemical Society. (e), (f) Reproduced with permission from the
American Chemical Society from Kuchibhatla et al.65 (g), (h) Reproduced from
Sayle et al.45 with permission from the Royal Society of Chemistry.

b1469_Ch-05.indd 282 4/8/2013 12:31:41 PM


b1469 Catalysis by Ceria and Related Materials

Atomistic Modelling of Ceria Nanostructures: Introducing Structural Complexity 283

various architectures, including nanochains, rods and superlattices,


as the crystallising ice front forces the nanoparticles together. Self-
assembly has been simulated at the atomistic level. In particular,
Fig. 5.15 shows model ceria nanoparticles self-assembled into
nanochains. Such models can be directly compared to experiment
(HRTEM and 3D tomography); a review of self-assembly is given in
Grzelczak et al.63
Similar to simulating the ‘self-assembly’ of ceria nanoparticles
into nanochains (1D assembly), the nanoparticles can also be assem-
bled into nanosheets (2D assembly) and framework architectures
(3D assembly).66 In particular, Fig. 5.16 shows spherical Ti–CeO2
core-shell nanoparticles assembled in three dimensions to form
mesoporous architectures. Under MD simulations the nanoparticles
attach and agglomerate together to form the walls of the mesoporous

{111}

CO

CO2

{100}

Figure 5.16 Assembly of ceria nanoparticles into mesoporous architectures. Top:


Schematic illustrating the principle. Bottom: Atomistic model of the mesoporous
ceria, which is enlarged (bottom right) to reveal the {111} and {100} planes exposed
at the internal surfaces of the material. Reprinted with permission from Sayle et al.46
Copyright 2007 American Chemical Society.

b1469_Ch-05.indd 283 4/8/2013 12:31:41 PM


b1469 Catalysis by Ceria and Related Materials

284 D. C. Sayle and T. X. T. Sayle

material. Simulated crystallisation was then employed to yield an


atomistic model of the mesoporous material with a microstructure
including point defects, dislocations and grain boundaries. The sys-
tem was then characterised using molecular graphics to help under-
stand the architecture and surfaces exposed. In particular, the {111}
and {100} surfaces were expressed at the wall surfaces.48 The wealth
of surface steps, edges, reactive {100} surfaces, curvature and high
surface area suggest that such mesoporous materials may have
exemplary catalytic properties. As with real mesoporous materials,
where characterisation is difficult, we note the integral part that
molecular graphics has to play in enabling one to characterise com-
plex atomistic models.

5.5.5.1 Systematic enumeration of nanostructure:


generating mesoporous models
One of the challenges for experimenters in synthesising mesoporous
materials is that one cannot position nanoparticles at any location
to engineer desired architectures; instead innovative ways of ena-
bling the nanoparticles to self-assemble into the desired positions
must be encouraged. On the other hand it is comparatively ‘easy’ to
construct an atomistic model using this strategy because model
nanoparticles can be placed in any position to formulate the nano-
structure — one ‘simply’ needs to type in the coordinates of where
one wishes the nanoparticles to be placed. In particular, Sayle and
co-workers explored the idea of using crystallographic ‘rules’ to
prescribe any nanostructure: analogous to crystal structures, by
simply changing the space symmetry and lattice positions of the
nanoparticles, one can generate a wealth of hypothetical nanostruc-
tures (superlattices) — this process provides a way of mapping
nanostructure.44 For example, in Fig. 5.17, nanoparticles of MgO
are positioned at FCC positions and agglomerate to form a frame-
work architecture.
The final models thus have various hierarchical levels of struc-
tural complexity (Fig. 5.18). They include the architecture of the
mesoporous material, the presence of microstructural features

b1469_Ch-05.indd 284 4/8/2013 12:31:42 PM


b1469 Catalysis by Ceria and Related Materials

Atomistic Modelling of Ceria Nanostructures: Introducing Structural Complexity 285

FCC

120o
CUBIC TETRAGONAL HEXAGONAL
a=b=c; α=β=γ=90o a=b≠c; α=β= γ=90o a=b≠c; α=β=90; γ=120o

[100] [110] [111]

Figure 5.17 Procedure for generating atomistic models for mesoporous materials
by positioning nanoparticles at crystallographic positions. The images at the bot-
tom are atomistic models of mesoporous MgO generated by positioning MgO
nanoparticles at FCC positions. Reprinted with permission from Sayle et al.44
Copyright 2008 American Chemical Society.

including dislocations, grain boundaries and point defects, and


finally the crystal structure and surfaces exposed owing to the orien-
tation of the crystal structure with respect to the surface and nano-
architecture. These structural models can be used to calculate
pertinent properties and processes. For catalysis, one can calculate
surface oxygen vacancy formation and simulate oxygen transport in
the material, which are pivotal to the catalytic properties of the
material.

b1469_Ch-05.indd 285 4/8/2013 12:31:42 PM


b1469 Catalysis by Ceria and Related Materials

286 D. C. Sayle and T. X. T. Sayle

Figure 5.18 Atomistic model for mesoporous MgO showing the three levels of
structural complexity: (left) crystal structure, (middle) microstructure and (right)
nanostructure. Reprinted with permission from Sayle et al.44 Copyright 2008
American Chemical Society.

Atomistic computer simulation can be used in a truly predictive


capacity to explore new porous architectures and elucidate their
properties. In particular, those architectures that are predicted
to proffer exemplary catalytic properties can be synthesised.
Accordingly, simulation can begin to be used to screen viable nano-
structures for important (catalytic) properties.

5.5.5.2 Mechanical properties


The operation of ceria under catalytic conditions can place the
material under severe mechanical duress and therefore it is impor-
tant to understand the behaviour of the material under operational
conditions, such as vibration, friction, thermal cycling, etc. The
mechanical properties of the material may prove pivotal. In particu-
lar, it is well known that microstructural features, such as disloca-
tions, defects and grain boundaries, govern the mechanical
properties and result in the measured ‘mechanical strength’ being
considerably lower than that predicted based upon the pristine,
defect-free material. If one is to simulate the mechanical properties
directly then atomistic models, which include all such microstruc-
tural features including their synergy of interaction, are needed.
And while there are considerable efforts focused in this direction,

b1469_Ch-05.indd 286 4/8/2013 12:31:44 PM


b1469 Catalysis by Ceria and Related Materials

Atomistic Modelling of Ceria Nanostructures: Introducing Structural Complexity 287

for example, see Sayle et al.47 and Sayle and Sayle,67 it is becoming
apparent that nanomaterials have the capability to proffer ‘ultra-
strength’ phenomena68 in that they can support stresses up to ‘a
significant fraction (>1/10) of their ideal strength — the highest
achievable stress of a defect-free crystal at zero temperature’. In par-
ticular, it has been predicted (via simulation) that ceria nanorods
can sustain yield strengths up to 50 GPa.47 Moreover, experimental
studies have revealed that ceria nanoparticles, 3 nm in diameter,
under quasihydrostatic pressure, are stable up to 60 GPa compared
to 30 GPa associated with the parent bulk material.69 It is likely
that the field of nanomechanics will unravel new physics pertain-
ing to nano(porous) materials with important implications for
nanocatalysis.

5.6 Ab Initio Methods


Atomistic simulation is well placed to calculate processes pertinent
to catalysis such as oxygen vacancy formation and oxygen mobility.
However, such simulations are critically dependent upon the intera-
tomic potentials employed to describe the material and cannot
address the electronic behaviour; instead ab initio methods are
required to explore processes that include bond breaking and bond
formation such as catalysis. For example, Nolan has explored oxy-
gen vacancy formation in ceria using hybrid density functional the-
ory (DFT),70 similarly, one can also calculate the oxidation process
CO + O = CO2 using oxygen extracted from a particular surface of
ceria.71 Typically, the results of these DFT calculations have revealed
the relative stability of the ceria surface and the ease of oxygen
extraction from them, in general accord with earlier atomistic simu-
lations. However, the computation cost associated with such quan-
tum mechanical (QM) techniques are high, which has prevented,
thus far, the inclusion of microstructure in the models. Indeed, as
alluded to above, the influence of this microstructure on catalytic
properties — specifically, the energetics and transport of oxygen in
the material — can be so profound that atomistic simulations can
prove more insightful than DFT calculations because the model

b1469_Ch-05.indd 287 4/8/2013 12:31:44 PM


b1469 Catalysis by Ceria and Related Materials

288 D. C. Sayle and T. X. T. Sayle

used for DFT calculations is necessarily more simplistic for the calcu-
lation to be tractable using current computational facilities.
Consequently, such methods are used in parallel with atomistic simu-
lation to provide complementary information for experiments.
To help circumvent this limitation, an approach to couple QM
with atomistic modelling, while at present in its infancy, has enjoyed
much success. Such quantum mechanics/molecular mechanics
(QM/MM) simulations include an inner core region, which com-
prises relatively few atoms and is treated quantum mechanically,
surrounded by a larger region which is treated atomistically,
Fig. 5.19. While the challenges associated with this method are con-
siderable–for example, the need to ensure the potential models
describing the MM component are commensurate with QM72 and
include dynamic reassignment ‘on-the-fly’ of the QM domain as the
‘reaction’ proceeds73–they promise the ability to simulate catalysis
explicitly spanning several hierarchical structural levels including:
electronic structure < crystal structure < microstructure < meso-
structure and ‘continuum’. In particular, the advantage of the
method is that they do not consider each hierarchical level in turn,
but similar to experiments, may be performed in a single simulation

CO CO2

Figure 5.19 QM/MM strategy. At the core is the QM region.

b1469_Ch-05.indd 288 4/8/2013 12:31:44 PM


b1469 Catalysis by Ceria and Related Materials

Atomistic Modelling of Ceria Nanostructures: Introducing Structural Complexity 289

thus capturing the synergistic interactions across the hierarchical


length scales spanning the electronic structure to a continuum.
As computational costs will inevitably decrease, the atomistic
simulations described here will likely be mirrored using quantum
mechanical methods.

5.7 Outlook
Atomistic computer simulation has continued to provide experi-
menters with unique insights and predictions. However, capturing
the hierarchical complexity associated with nanomaterials, within a
single atomistic model, is difficult; perhaps the easiest way to gener-
ate such models is by simulating, in part, the synthetic method used
during their manufacture. Moreover, a benefit of this approach is
the ability to be able to make direct comparisons between experi-
ment and simulation.
On the other hand, if the models comprising all hierarchical
levels of structural complexity are generated by ‘simulating synthe-
sis’ rather than generating each component ‘by-hand’ using symme-
try operators, one loses ultimate control over the atomistic structure;
instead, similar to experiments, control over the microstructures
and architectures may only be achieved indirectly by controlling, for
example, temperature, pressure and methods of (self-) assembly.
QM/MM methods74,75–which enable one to simulate the core
region quantum mechanically, to model bond formation or break-
ing, coupled with an atomistic level description of the surrounding
regions, which include microstructural features and nanoarchitec-
ture, together with continuum methods that effectively extend the
material to ‘infinity’–while at time of writing are in their infancy,
span hierarchical length scales from the electronic structure to a
continuum and have the potential to truly ‘Simulate the dirt’!

Acknowledgements
The authors gratefully acknowledge funding from EPSRC EP/
H001220.

b1469_Ch-05.indd 289 4/8/2013 12:31:45 PM


b1469 Catalysis by Ceria and Related Materials

290 D. C. Sayle and T. X. T. Sayle

References
1. Catlow, C.R.A., Smith, B.J., van Santen, R.A., eds., Computer Modelling
of Microporous Materials, Academic Press, Amsterdam, (2004).
2. Smith, W., Forester, T.R., DL_POLY, copyright by the Council for the
Central Laboratory of the Research Councils, Daresbury Laboratory,
Daresbury, Warrington, UK (1996). www.cse.clrc.ac.uk/msi/software/
DL_POLY/.
3. Gale, J.D., Rohl, A.L., Mol. Simulat., 29 (2003) 291–341.
4. Hamad, S., Catlow, C.R.A., Woodley, S.M., Lago, S., Mejías, J.A.,
J. Phys. Chem. B, 109 (2005) 15741–15748.
5. Johnston, R.L., Dalton T., (2003) 4193–4207.
6. Sayle, T.X.T., Parker, S.C., Catlow, C.R.A., Chem. Commun., 14 (1992)
977–978.
7. Balducci, G., Islam, M.S., Kaspar, J., Fornasiero, P., Graziani, M., Chem.
Mater., 15 (2003) 3781–3785.
8. Goniakowski, J., Finocchi, F., Noguera, C., Rep. Prog. Phys., 71 (2008)
016501.
9. Watson, G.W., Kelsey, E.T., de Leeuw, N.H., Harris, D.J., Parker, S.C.,
J. Chem. Soc., Faraday T., 92 (1996) 433–438.
10. Norenberg, H., Harding, J.H., Surf. Sci., 477 (2001) 17–24.
11. Conesa, J.C., Surf. Sci., 339 (1995) 337–352.
12. Mai, H.X., Sun, L.D., Zhang, Y.W., Si, R., Feng, W., Zhang, H.P.,
Liu, H., Yan, C.H., J. Phys. Chem. B, 109 (2005) 24380–24385.
13. Niesz, K., Reji, C., Neilson, J.R. Vargas, R.C., Morse, D.E., ACS Crystal
Growth and Design, 10 (2010) 4485–4490.
14. Zhang, J., Kumagai, H., Yamamura, K., Ohara, S., Takami, S.,
Morikawa, A., Shinjoh, H., Kaneko, K., Adschiri, T., Suda, A., Nano
Lett., 11 (2011) 361–364.
15. Wu, Z., Li, M., Howe, H., Meyer, H.M., Overbury, S.H., Langmuir, 26
(2010) 16595–16606.
16. Gritschneder, S., Reichling, M., Nanotechnology, 18 (2007) 044024.
17. Gritschneder, S., Namba, Y., Reichling, M., Nanotechnology, 16 (2005)
S41–S48.
18. Baudin, M., Wojcik, M., Hermansson, K., Palmqvist, A.E.C.,
Muhammed, M., Chem. Phys. Lett., 335 (2001) 517–523.

b1469_Ch-05.indd 290 4/8/2013 12:31:45 PM


b1469 Catalysis by Ceria and Related Materials

Atomistic Modelling of Ceria Nanostructures: Introducing Structural Complexity 291

19. Li, H.-Y., Wang, H.-F., Guo, Y.-L., Lu, G.-Z., Hu, P., Chem. Commun., 47
(2011) 6105–6107.
20. Gotte, A., Hermansson, K., Baudin, M., Surf. Sci., 552 (2004) 273–280.
21. Sayle. D.C., J. Mater. Chem., 9 (1999) 2961–2964.
22. Zhang, F., Walker, A.M., Wright, K., Gale, J.D., J. Mater. Chem., 20
(2010) 10445–10451.
23. Hojo, H., Mizoguchi, T., Ohta, H., Findlay, S.D., Shibata, N.,
Yamamoto, T., Ikuhara, Y., Nano Lett., 10 (2010) 4668–4672.
24. Tasker, P.W., J. Chem. Soc. Faraday T., 86 (1990) 1311–1315.
25. Duffy, D.M., Tasker, P.W., Philos. Mag. A, 54 (1986) 759–771.
26. Fisher, C.A.J., Matsubara, H., J. Euro. Ceram. Soc., 19 (1999) 703–707.
27. Sutton, A.P., Balluffi, R.W., Interfaces in Crystalline Materials: Monographs
on the Physics and Chemistry of Materials, Clarendon Press, Oxford,
(1995).
28. Sayle, T.X.T., Catlow, C.R.A., Sayle, D.C. Parker, S.C., Harding, J.H.,
Philos. Mag. A, 68 (1993) 565–573.
29. Sayle, D.C., Maicaneanu, A.S. Watson, G.W., J. Am. Chem. Soc., 124
(2002) 11429–11439.
30. Martin, P., Parker, S.C., Sayle, D.C., Watson, G.W., Nano Lett., 7 (2007)
543–546.
31. Zhou, K., Yang, Z., Yang, S., Chem. Mater., 19 (2007) 1215–1217.
32. Cordatos, H., Ford, D., Gorte, R.J., J. Phys. Chem., 100 (1996)
18128–18132.
33. Sayle, D.C., Maicaneanu, S.A., Slater, B., Catlow, C.R.A., J. Mater.
Chem., 9 (1999) 2779–2787.
34. Mohn, C.E., Stein, M.J., Allan, N.L., J. Mater. Chem., 20 (2010)
10403–10411.
35. Piana, S., Reyhani, M., Gale, J.D., Nature, 438 (2005) 70–73.
36. Phillpot, S.R., Keblinski, P., Wolf, D., Cleri, F., Interface Sci., 7 (1999)
15–31.
37. Sayle, D.C., Watson, G.W., J. Phys. Chem. B, 106 (2002) 3778–3787.
38. Sayle, D.C., Watson, G.W., J. Mater. Chem., 10 (2000) 2241–2243.
39. Quigley, D., Rodger, P.M., Mol. Simulat., 35 (2009) 613–623.
40. Anwar, J., Zahn, D., Angew. Chem. Int. Ed., 50 (2011) 1996–2013.
41. Streitz, F.H., Glosli, J.N., Patel, M.V., Phys. Rev. Letters, 96 (2006)
225701.

b1469_Ch-05.indd 291 4/8/2013 12:31:45 PM


b1469 Catalysis by Ceria and Related Materials

292 D. C. Sayle and T. X. T. Sayle

42. Sayle, T.X.T., Parker, S.C., Sayle, D.C., Chem. Commun., 21 (2004)
2438–2439.
43. Feng, X.D., Sayle, D.C., Wang, Z.L., Paras, M.S., Santora, B., Sutorik,
A., Sayle, T.X.T., Yang, Y., Ding, Y., Wang, X.D., Her, Y.S., Science, 312
(2006) 1504–1508.
44. Sayle, D.C., Seal, S., Wang, Z.W., Mangili, B.C., Price, D.W.,
Karakoti, A.S., Kuchibhatla, S.V.T.N., Hao, Q., Moebus, G., Xu, X.,
Sayle, T.X.T., ACS Nano, 2 (2008) 1237–1251.
45. Sayle, T.X.T., Parker, S.C., Sayle, D.C., Faraday Discuss., 134 (2007)
377–397.
46. Sayle, D.C., Feng, X., Ding, Y., Wang, Z.L., Sayle, T.X.T., J. Am. Chem.
Soc., 129 (2007) 7924–7935.
47. Sayle, T.X.T., Inkson, B.J., Karakoti, A., Kumar, A., Molinari, M.,
Mobus, G., Parker, S.C., Seal, S., Sayle, D.C., Nanoscale, 3 (2011)
18223–18237.
48. Du, N., Zhang, H., Chen, B., Ma, X., Yang, D., J. Phys. Chem. C, 111
(2007) 12677–12680.
49. Barnard, A.S., Rep. Prog. Phys., 73 (2010) 086502.
50. Catlow, C.R.A., Stefan, T., Brombley, T., Hamad, S., Mora-Fonz, M.,
Sokola, A., Woodley, S.M., Phys. Chem. Chem. Phys., 12 (2010)
786–811.
51. Sayle, T.X.T., Parker, S.C., Sayle, D.C., Phys. Chem. Chem. Phys., 7
(2005) 2936–2941.
52. Migani, A., Vayssilov, G.N., Bromley, S.T., Illas, F., Neyman, K.M.,
J. Mater. Chem., 20 (2010) 10535–10546.
53. Liu, X., Zhou, K., Wang, L., Wang, B., Li, Y, J. Am. Chem. Soc., 131
(2009) 3140–3141.
54. Sata, N., Eberman, K., Eberl, K., Maier, J., Nature, 408 (2000) 946–949.
55. Sayle, D.C., Doig, J.A., Parker, S.C., Watson, G.W., Sayle, T.X.T., Phys.
Chem. Chem. Phys., 7 (2005) 16–18.
56. Sayle, T.X.T., Parker, S.C., Sayle, D.C., J. Mater. Chem., 16 (2006)
1067–1081.
57. Ji, P., Zhang, J., Chen, F., Anpo, M., J. Phys. Chem. C, 112 (2008)
17809–17813.
58. Möbus, G., Inkson, B.J., Materials Today, 10 (2007) 18–25.

b1469_Ch-05.indd 292 4/8/2013 12:31:45 PM


b1469 Catalysis by Ceria and Related Materials

Atomistic Modelling of Ceria Nanostructures: Introducing Structural Complexity 293

59. Satyanarayana, V.N., Kuchibhatla, A.S., Karakoti, S., Bera, D., Seal, S.,
Prog. Mat. Sci., 52 (2007) 699–913.
60. Ji, P., Zhang, J., Chen, F., Anpo, M., J. Phys. Chem. C, 112 (2008)
17809–17813.
61. Smarsly, B., Antoinietti, M., Eur. J. Inorg. Chem., (2006) 1111–1119.
62. Deville, S., Saiz, E., Nalla, R.K., Tomsia, A.P., Science, 311 (2006)
515–518.
63. Grzelczak, M., Vermant, J., Furst, E.M., Liz-Marza, L.M., ACS Nano, 4
(2010) 3591–3605.
64. Tan, J.P.Y., Tan, H.R., Boothroyd, C., Foo, Y.L, He, C.B., Lin, M.,
J. Phys. Chem. C, 115 (2011) 3544–3551.
65. Kuchibhatla, S.V.N.T., Karakoti, A.S., Sayle, D.C., Heinrich, H.,
Seal, S., Cryst.Growth Des., 9 (2009) 1614–1620.
66. Sayle, T.X.T., Maphanga, R.R., Ngoepe, P.E., Sayle, D.C., J. Am. Chem.
Soc., 131 (2009) 6161–6173.
67. Sayle, T.X.T., Sayle, D.C., ACS Nano, 4 (2010) 879–886.
68. Zhu, T., Li, J., Prog. Mater. Sci., 55 (2010) 710–757.
69. Wang, Z.W., Seal, S., Patil, S., Zha, C.S., Xue, Q., J. Phys. Chem. C, 111
(2007) 11756–11759.
70. Nolan, M., Chem. Phys. Lett., 499 (2010) 126–130.
71. Nolan, M., Parker, S.C., Watson, G.W., Phys. Chem. Chem. Phys., 8
(2006) 216–218.
72. Galea, N.M., Scanlon, D.O., Martin, P., Watson, G.M., Sherwood, P.,
J. Surf. Sci. Nanotechnol., 7 (2009) 413–420.
73. Kermode, J.R., Albaret, T., Sherman, D., Bersteinet, N., Gumbsch, P.,
Payne, M.C., Csanyi, G., De Vita, A., Nature, 455 (2008) 1224–1227.
74. Sherwood, P., de Viers, A.H., Guest, M.F., Schreckenbach, G.,
Catlow, C.R.A., French, S.A., Sokol, A.A., Brombley, S.T., Thiel, W.,
Turner, A.J., Billeter, S., Terstegen, F., Thiel, S., Kendrick, J.,
Rogers, S.C., Casci, J., Watson, M., King, F., Karlsen, E., Sjovoll, M.,
Fahmi, A., Scheafer, A., Lennartz, C., J. Mol. Struct., 632 (2003) 1–28.
75. Sokol, A.A., French, S.A., Bromley, S.T. Catlow, C.R.A., van Dam, H.J.J.,
Sherwood, P., Faraday Discuss., 134 (2007) 267–282.
76. Sankaranarayanan, S.K.R..S., Ramanathan, S., J.Chem. Phys., 134 (2011)
064703.

b1469_Ch-05.indd 293 4/8/2013 12:31:45 PM


b1469 Catalysis by Ceria and Related Materials

This page intentionally left blank

b1469_Ch-05.indd 294 4/8/2013 12:31:45 PM


b1469 Catalysis by Ceria and Related Materials

CHAPTER 6

TWO-DIMENSIONAL AND
THREE-DIMENSIONAL CERIA-BASED
NANOARCHITECTURES
Zhen-Xing Li, Wei Feng, Chao Zhang, Ling-Dong Sun,
Ya-Wen Zhang, and Chun-Hua Yan
Beijing National Laboratory for Molecular Science, State Key Laboratory
of Rare Earth Materials Chemistry and Applications, PKU-HKU
Joint Laboratory on Rare Earth Materials and Bioinorganic Chemistry,
Peking University, Beijing, 100871, China

6.1 Introduction
Reducing the size of inorganic materials to the nanometer scale has
been one of the most attractive realms in materials science.
Nanomaterials are notable for their featured small size, which
endows them with size- and shape-dependent physical and chemical
properties owing to the high surface/volume ratio and quantum
effects.1–10 With the booming evolution of this field, different nano-
materials have been developed in several large families.11–20 Ceria-
based materials are of great interest for researchers all over the
world due to their many applications, in particular as redox or oxy-
gen storage promoters in three-way catalysts,21–23 catalysts for H2
production from fuels, and solid-state conductors for fuel cells.24–27
Pure stoichiometric CeO2 has a calcium fluoride (fluorite)-type
structure with space group Fm3m over the whole range from room
temperature to its melting point; its color is pale yellow, probably
due to Ce(IV)–O(II) charge transfer. Ceria, in the fluorite structure,

295

b1469_Ch-06.indd 295 4/8/2013 12:32:31 PM


b1469 Catalysis by Ceria and Related Materials

296 Z.-X. Li et al.

exhibits defects depending on the partial pressure of oxygen, which


is the intrinsic property for its potential in catalysis, energy conver-
sion, and other fields.
Nanoscience and nanotechnology may lead to many new appli-
cations for these materials. As with most nanomaterials, as the parti-
cle size of ceria decreases down to the nanometer scale its bandgap
increases, reflected by the blue shift in the absorption spectra owing
to the quantum size effect.28 The existence of Ce3+ ions in ceria par-
ticles as well as the oxidation state as a function of particle size was
investigated using X-ray photoelectron spectroscopy29 and electron
energy loss spectroscopy,30 respectively. In most of the envisaged
applications, catalytic performance is strongly dependent on particle
size; for instance, large conventional ceria particles are not good
candidates as catalysis supports for CO oxidation, whereas ceria
nanoparticles (NPs) show high catalytic activity.31 For electrical
applications, ceria has also been observed to show grain size-
dependent electrical conductivity.32 Recently, besides the conven-
tional application fields mentioned above, ceria-based nanomaterials
have been used in nanodevices.33 All these features make ceria-based
nanomaterials very attractive for chemists.
This chapter attempts to highlight current research activities
centered on the controlled synthesis and assembly of ceria-based
nanomaterials. We also introduce our own endeavors on this theme
over the past ten years and give a general outlook. This chapter
briefly reviews solution-based synthetic routes and the catalytic prop-
erties of ceria-related nanomaterials developed in the last decade.
The first part covers the recently developed synthetic routes for
ceria nanomaterials, mainly concerning the shape control of
nanocrystals to obtain nanoparticles (0D), nanorods and nanowires
(1D), and aggregates and mesostructures (2D or 3D). The second
part of the chapter reviews theoretical and experimental results con-
cerning the relation between catalytic properties and the structure
of ceria nanomaterials. The fabrication of ceria–metal nanocompos-
ites and the corresponding catalytic applications are also included.
Finally, we summarize the review and provide a brief outlook for this
exciting field.

b1469_Ch-06.indd 296 4/8/2013 12:32:32 PM


b1469 Catalysis by Ceria and Related Materials

2D and 3D Ceria-Based Nanoarchitectures 297

6.2 Synthesis and Assembly of Ceria-Related Nanomaterials


6.2.1 Synthesis of ceria-based nanomaterials
Ceria has a cubic phase with a fluorite structure. Therefore, common
as-prepared ceria nanocrystals are isotropic in shape. On the other
hand, the fluorite structure has a different atomic arrangement in
different facets, which leads to distinct catalytic activity. Controlling
the shape of nanocrystals can be used to tune the catalytic properties
of such materials. The size of nanocrystals is another important factor
in determining catalytic properties because of the surface effect.34
Various methods have been developed to control the size and shape
of ceria nanomaterials. In this section, we will describe recent advances
in the synthesis of this kind of nanomaterial and present some typical
examples to introduce the characteristics of the various methods.

6.2.1.1 Zero-dimensional (0D) ceria nanoparticles: nanopolyhedra


As described in the introduction, 0D ceria nanoparticles are easy to
obtain because the cubic phase structure has no special favorite
growth direction.
The common synthesis techniques of nanocrystals can be easily
applied to the synthesis of ceria nanocrystals. Because of the ultra-low
solubility of ceria, basic co-precipitation can be applied directly and
commercial ceria nanocrystals are obtained by this simple method.
An alkali solution is added to a Ce(III) or Ce(IV) solution to initiate
the hydrolysis of Ce(III) and Ce(IV) ions. Ce(OH)3 can be easily
oxidized to Ce(OH)4. The as-produced Ce(OH)4 quickly loses H2O
and forms ceria. Because of its ultra-low solubility, ceria nanocrystals
are usually small and hard to grow, which is different from the usual
recrystallization in the synthesis of other kinds of nanocrystals.
Modified co-precipitation methods are also applied in the syn-
thesis of ceria nanocrystals. NaOH,35,36 ammonia,37,38 urea,39,40 and
Na2CO341 can all be used as precipitation agents. Cetyl trimethylam-
monium bromide (CTAB),42 carboxymethylcellulose sodium,43 and
polyvinylpyrrolidone (PVP)44,45 are used as surfactants to control the
size of the nanocrystals.

b1469_Ch-06.indd 297 4/8/2013 12:32:32 PM


b1469 Catalysis by Ceria and Related Materials

298 Z.-X. Li et al.

Figure 6.1 Typical high resolution transmission electron microscopy (HRTEM)


images of CeO2 nanoparticles with irregular shapes. The inset in (b) is a Fourier
transform of the corresponding HRTEM image of one irregular shape nanoparticle.
Reprinted with permission from Wang and Feng.53 Copyright 2003 the American
Chemical Society.

Besides the co-precipitation method, other common methods


for the synthesis of nanomaterials, such as the hydrothermal
method,46–53 the solvothermal method,54 the sonochemical method,55
pyrolysis,56 the sol-gel process,57,58 and the reverse micelles method
have also been applied for the synthesis of ceria nanomaterials. The
typical shape of as-prepared nanocrystals is always polyhedral
because of the fluorite crystal structure of ceria (Fig. 6.1).

b1469_Ch-06.indd 298 4/8/2013 12:32:32 PM


b1469 Catalysis by Ceria and Related Materials

2D and 3D Ceria-Based Nanoarchitectures 299

Oh et al. investigated the influence of the dielectric properties of


solvents used in the hydrothermal reaction on the morphology and
particle size of ceria nanocrystals.59 Cerium nitrate and KOH were
used as starting materials, and methanol, ethanol, ethylene glycol,
and 1,4-butylene glycol were chosen as different dielectric solvents.
They found that the size of the final products increased along with
an increase of the dielectric constant of the solvent and reaction tem-
perature, and a decrease in the pH values of the reaction medium.
Li and co-workers investigated the evolution of the morphology
in the solvothermal synthesis of ceria nanoparticles.60 Rod-like and
prism-like Ce(OH)CO3 and polygon-like CeO2 can be obtained with
urea as precipitator. Using the organic base triethanolamine or inor-
ganic NaOH would result in cube-like ceria nanocrystals.
Adachi and co-workers synthesized ultra-fine ceria nanocrystals
within reverse micelles.61 The particle size was less than 5 nm, which
can be tuned by applying different water/oil ratios.
Ceria nanoparticles can also be synthesized in organic solvents.
Yan and co-workers used cerium alkoxide as a precursor to be
decomposed in a high-boiling solvent (tri-n-octylphosphine oxide
and dioctyl ether).62 The structure of the final product can be tuned
by changing reaction conditions such as temperature and the
concentration of the precursors.
Recently, ionic liquids have been also employed as solvents to syn-
thesize ceria nanoparticles. Goharshadi et al. used a set of ionic liquids
based on the bis(trifluoromethylsulfonyl)imide anion and different
cations of 1-alkyl-3-methyl-imidazolium.63 After microwave-assisted
heating, ceria nanoparticles were synthesized by precipitation.
Another method of restricting the size of the nanocrystals is to
employ a template. Wilson et al. introduced the use of pyrolyzing ion
exchange resins loaded with cerium to obtain ceria nanoparticles
within carbon support matrixes.64 The as-prepared hybrid structures
contained sub-2 nm ceria nanoparticles with a high surface area.
Because of their different surface structures and catalytic prop-
erties, the exposed facets of the ceria nanoparticles also need to be
tuned.65–67 The most common exposed facets of ceria nanocrystals
are {200}, {111}, and {110}. Both theoretical and experimental results

b1469_Ch-06.indd 299 4/8/2013 12:32:32 PM


b1469 Catalysis by Ceria and Related Materials

300 Z.-X. Li et al.

Figure 6.2 Modeling the atom positions of a single nanoparticle using molecular
dynamics through an annealing process. Top, starting configurations, left to right:
(a): {100}; (b): {110}; (c): {111} and {110}. Middle: final configurations. Bottom: final
configurations with surface-filling representation to aid interpretation. Cerium is
white and oxygen is black. Reprinted with permission from Sayle et al.70 Copyright
2004 the Royal Chemical Society.

have proved that the surface energies of these facets are in the fol-
lowing order: {111} < {200} < {110} in vacuum.53,68,69 So the common
exposed facets are {111} and {200}, which result in the polyhedral
({111} + {200}), octahedral (pure {111}), or cubic (pure {200}) shape
of the final product. Because the surface energies of {111} and {200}
are nearly the same in ambient conditions, the final products are
usually polyhedra (Fig. 6.2).53,70
Different anions are applied in the synthetic procedures to
adjust the surface energies of different facets of ceria, and therefore

b1469_Ch-06.indd 300 4/8/2013 12:32:32 PM


b1469 Catalysis by Ceria and Related Materials

2D and 3D Ceria-Based Nanoarchitectures 301

Figure 6.3 (a)–(c) TEM images of ceria nanocubes with average sizes of (a) 4.43 nm,
(b) 7.76 nm, and (c) 15.65 nm; the insets are selected-area electron diffraction
(SAED) patterns of individual NPs. (d) High-resolution transmission electron
microscopy (HRTEM) image of ceria nanocubes. (e) Typical X-ray diffraction pat-
tern of ceria nanocubes assembled on an Si wafer; the inset is the schematic illustra-
tion of the facets of an individual cube. Reprinted with permission from Yang and
Gao.71 Copyright 2006 the American Chemical Society.

influence the final shape of the nanomaterials. Gao and co-workers


used oleic acid as a surface stabilization agent and obtained uniform
small nanocubes (Fig. 6.3).71 Taniguchi et al. used oleate as a capping
ligand in the hydrothermal reaction to synthesize nanocubes.72 The
size of the nanocubes can be tuned by using different Ce precursors.
Wu et al. investigated a series of anions to mediate the growth of
ceria nanocrystals with the hydrothermal reaction.73 They found
that NO3− leads to nanocubes, while Cl−, Br−, I−, and SO42− anions
lead to the formation of nanorods. Kaneko et al. used decanoic acid
as a capping reagent in the hydrothermal reaction to obtain ceria
nanocubes.234 They also performed theoretical simulations for the
high-resolution (HR) TEM images to determine the atomic

b1469_Ch-06.indd 301 4/8/2013 12:32:33 PM


b1469 Catalysis by Ceria and Related Materials

302 Z.-X. Li et al.

arrangement and terminating atoms. The evolution of the shape is


clearly shown by the simulation results.
Chen, Yang, and co-workers investigated the Yb3+ doping effect to
control the size and shape of ceria nanocrystals.74 Doped by trivalent
rare-earth ions, the surfaces of the nanocrystals adsorb more OH−,
which accelerates recrystallization to give regular-shaped nanocubes.

6.2.1.2 One-dimensional (1D) nanomaterials: nanorods, nanowires,


nanotubes, and nanospindles
One-dimensional nanostructures have attracted attention because
of their special properties and assembly characteristics. The general
method of producing 1D nanomaterials has been reviewed.20 The
synthesis of 1D ceria nanoparticles has also recently been reviewed,75
so this section will only give a brief introduction to the various meth-
ods of producing these structures.
Generally, 1D ceria nanocrystals can be obtained through two
methods. The first is the template method, which uses a hard or soft
template to restrict the growth of the ceria nanocrystals to one
specific direction. The other method adjusts surface energies in
which some facets are protected by specific capping ligands, which
restrict growth.
A hard template can restrict the growth of nanomaterials by
spatial confinement. To grow 1D nanomaterials, anodic aluminum
oxide (AAO) is a commonly used hard template with uniform hex-
agonally ordered channels. Assisted by the capillary driving force,
the precursors diffuse into the channel. After precipitation by a base
or simple hydrolysis by heating, the 1D nanostructure can form.76
Xia and co-workers described an AAO-directed self-combustion
method to obtain ceria nanotubes.77 Ethyl glycol, citric acid, and
metal nitrates were dissolved in solution and filled the channels of
the AAO template. The combustion reaction created ceria nano-
tubes within the channels.
Carbon nanotube (CNT) is another kind of hard template. After
surface modification, CNTs can adsorb OH− or Ce(III, IV) ions at
the surface, which result in 1D ceria nanostructures after precipita-
tion and calcination to remove the template.

b1469_Ch-06.indd 302 4/8/2013 12:32:33 PM


b1469 Catalysis by Ceria and Related Materials

2D and 3D Ceria-Based Nanoarchitectures 303

Li et al. used infrared (IR) spectra to investigate the surface func-


tional groups on the treated CNTs.78 After boiling with an oxidizing
acid, the peaks corresponding to the carbonyl and hydroxyl groups
are sharper. Zhang, Shi, and co-workers found that the zeta poten-
tial of CNTs was higher after treatment with nitric acid, which
proved that there were more acidic groups on the treated CNTs.79
These acidic groups adsorbed Ce3+ ions from the aqueous suspen-
sion and finally formed CeO2 nanotubes. Following this acid-etching
assisted method, there have been several reports of a similar synthe-
sis of ceria nanotubes.80–84
Song, Guo, and co-workers also used CNTs as a template to
obtain oxide nanotubes.85 They treated CNTs with glucose to form a
carbonaceous coating on the outside of the CNTs. The adsorption
of metal ions directed the growth of the oxide nanocrystals along
the CNTs. After calcination, oxide nanotubes were produced. They
prepared hollow structures of SnO2, ZrO2, CeO2, and Fe2O3 to prove
that this is a general method.
Chen et al. employed 1D attapulgite with a uniform diameter
ranging from 20 to 40 nm as a template to grow ceria nanotubes.86
Attapulgite rods were first treated with acid, followed by the precipi-
tation of ceria nanocrystals on the surface. After etching by an NaOH
solution, the template was removed from the ceria nanotubes.
Boehme et al. employed polycarbonate membrane filters as tem-
plates to synthesize ceria nanotubes by low-temperature deposi-
tion.87 The polycarbonate membranes were chemically etched by an
NaOH solution, followed by an activation and sensitization process
performed by aqueous SnCl2, AgNO3 and Co(NO3)2, Pd(NO3)2, and
Ag2SO4. The final electroless deposition step used a Ce(III) and
C2H10BN solution. Ceria nanotubes grew in the membrane channel
and the template was easily removed by dichloromethane.
Micelle systems are usually used as soft templates for the synthe-
sis of nanomaterials.
Yuan, Su, and co-workers used CTAB as a surfactant to modify
the morphology of nanocrystals.88 The authors realized that hydrated
Ce(IV) species can incorporate with CTA+ at some pH value and
together form a micelle structure to direct the growth of ceria
nanocrystals.

b1469_Ch-06.indd 303 4/8/2013 12:32:33 PM


b1469 Catalysis by Ceria and Related Materials

304 Z.-X. Li et al.

Jin and co-workers obtained 1.5-nm ultra-thin ceria nanorods from


a cerium stearate precursor.89 The reaction was carried up at 270°C in
1-octadecene with the addition of 1-octadecanol. The authors ascribed
the formation of nanorods to the oriented attachment of ultra-small
ceria nanocrystals grown in the initial stage of the reaction.
Another soft template system was used by Hyeon and co-workers.90
The synthesis was performed in an oleic acid–oleylamine system.
Nanoparticles and ultra-thin nanowires were obtained by tuning the
reaction conditions. The formation mechanism was considered to be
the oriented attachment of small ceria nanoparticles. This work is simi-
lar to recent work on the synthesis of ultra-thin gold nanowires.91,92
Template-free methods of producing 1D ceria nanomaterials
usually have intermediate products with 1D morphology or undergo
Ostwald ripening. As mentioned for the synthesis of 0D ceria nano-
particles, crystallization is difficult in the common precipitation of
ceria materials. Thus the Ostwald ripening usually took quite a long
time and was carried up under extreme conditions.
Han et al. described the facile synthesis of ceria nanotubes via
the precipitation of Ce(III) by ammonia hydroxide followed by an
aging process that lasted for 45 days (Fig. 6.4).93

Figure 6.4 (a) Typical morphology of ceria samples produced by a hydrothermal


route followed by an aging process. Nanoparticles, nanowires, and nanotubes are
marked. (b) High-resolution image of a nanowire. (c) High-resolution image of a
nanotube. Reprinted with permission from Han et al.93 Copyright 2005 the American
Chemical Society.

b1469_Ch-06.indd 304 4/8/2013 12:32:33 PM


b1469 Catalysis by Ceria and Related Materials

2D and 3D Ceria-Based Nanoarchitectures 305

Chen et al. described a two-step co-precipitation method to


obtain 1D ceria nanomaterials.94 The intermediates were a mixture
of Ce(III)–Ce(IV) oxides. They found that the aging temperature
can influence the final shape of the products.
Yan and co-workers used Ce(III) precursors to react with NaOH
using the hydrothermal method.95 The mechanism was considered
to be that hexagonal phase Ce(OH)3 nanorods were initially formed
in the reaction, followed by oxidization to the final ceria products
with a rod-shape.
Pan et al. used a similar hydrothermal method to prepare ceria
nanorods, nanotubes, and nanowires using a Ce(OH)3 intermedi-
ate.96 They investigated the influence of reaction temperature and
aging time, and explained the selective formation of nanowires
and nanotubes as due to Ostwald ripening. Wang et al. investigated
the effect of anions, such as Cl−, NO3−, and PO43−, on the morphology
of the Ce(OH)3 intermediates and the final products.97 Ceria
nanorods and nanowires were selectively obtained by tuning the
reaction conditions.
Ding, Sun, and co-workers used urea as a precipitant to obtain
Ce(OH)CO3 nanorods as intermediate products.98 After treatment
with NaOH at different concentrations, two kinds of ceria nano-
tubes were obtained through an interface reaction involving
Ce(OH)CO3 and Ce(OH)3 as intermediates, respectively. These
methods for synthesizing ceria nanotubes are a result of the
Kirkendall effect, involving the dissolution and oxidization of
Ce(III) followed by precipitation of Ce(IV).99
Zhu and co-workers described the facile ultrasonic-induced syn-
thesis of nanotubes.100 Ceria nanoparticles of 2.6 nm were used as
precursors. The reaction was carried in an alkali (10 M KOH) exposed
to high-intensity ultrasound. The ultrasound caused localized high
temperatures and pressures, which accelerated recrystallization.
Seal and co-workers described the similar growth of ceria
nanotubes in a double microemulsion system composed of bis(2-
ethylhexyl) sulfosuccinate sodium salt (AOT), toluene, and water.101
Ceria nanoparticles were formed inside the micelles and aggregated
into tubular particles after aging for a few weeks. This kind of

b1469_Ch-06.indd 305 4/8/2013 12:32:33 PM


b1469 Catalysis by Ceria and Related Materials

306 Z.-X. Li et al.

protocol can produce 1D ceria nanoparticles composed of ultra-


small nanocrystals, thus the products usually have large surface/
volume ratios and high catalytic activity.102
Liu and co-workers extended this precipitation and recrystalliza-
tion method to non-aqueous solutions.103 Supercritical CO2/ethanol
was used as a solvent with high diffusivity and strong salvation strength,
which are good for the recrystallization of 1D nanostructures.
Sayle and co-workers proposed an atomistic model of ceria
nanotubes and predicted that this kind of structure will further
unlock the oxygen storage capacity (OSC) of the materials, which
will benefit the catalytic properties.104
Another kind of 1D ceria nanostructure is the spindle. Zheng and
co-workers described a hydrothermal method to obtain spindles.105
Urea was used as a precipitant and the reaction time was relatively
shorter than for nanotubes. Zhang, Shi and co-workers also described
a method to prepare ceria nanospindles. Ce(III) was precipitated by
urea in the presence of glycerin. Ce(OH)CO3 was realized as interme-
diate and the mechanism involved the assembly of small particles
assisted by glycerin molecules adsorbed on the surface of the parti-
cles.106 Sun et al. used a similar solvothermal method to assemble
shuttle-like ceria nanomaterials from nanorods.107 Zhang, Tong and
co-workers described the synthesis of bowknot-like ceria bundles
assembled from nanorods obtained via a hydrothermal reaction.108

6.2.1.3 Two-dimensional (2D) nanomaterials: nanoplates and nanodisks


Two-dimensional ceria nanocrystals are usually obtained through
the template method because it is difficult to tune the growth rate
of different facets to form this type of morphology.
Soft templates have also been used for the synthesis of ceria
nanoplates. Murray and co-workers used a high-boiling solvent
system for the synthesis of ceria nanocrystals.109 Oleic acid and
oleylamine were chosen as the solvent and capping agents and
sodium diphosphate or sodium oleate were used as mineralizers in
the experiment. Square-shaped ceria nanoplates were prepared and
assembled into wire-like structures.

b1469_Ch-06.indd 306 4/8/2013 12:32:33 PM


b1469 Catalysis by Ceria and Related Materials

2D and 3D Ceria-Based Nanoarchitectures 307

Figure 6.5 Scanning electron microscopy (SEM) images of (a) plate-like precur-
sors and (b) final ceria products after calcining at 400°C. Reprinted with permis-
sion from Minamidate et al.110 Copyright 2010 Elsevier B.V.

Yin and co-workers used NaHCO3 and Na2CO3 as precipitants to


prepare plate-like cerium carbonate precursors (Fig. 6.5).110 After
calcination at 400°C for 1 hour, ceria particles with the right shape
were obtained. The particle size can be tuned by adjusting the pH
value of the solution during precipitation.
Zhong and co-workers described a template-free method to syn-
thesize ceria microdisks.111 Self-assembled hierarchical Ce(SO4)2
microdisks were synthesized via a solvothermal method as an inter-
mediate product. After calcination at 1100°C, ceria microdisks were
obtained.
Zhang, Shi, and co-workers described a method to obtain rhom-
bic microplates.112 A simple water reflux method was employed for
the precipitation of ceria by urea with the assistance of CTAB. The
authors also realized that the microdisks evolved from small ceria
nanoparticles.
Another method to obtain 2D ceria nanocrystals is the electro-
chemical route. Li, Tong, and co-workers developed the method,
which involves the electrochemical deposition of Ce(III) in a
NH4NO3 solution.113 Ceria nanosheets and nanobelts were obtained
by applying lower and higher current densities, respectively. The
CeO2 and Ce(OH)3 nuclei were proposed as two important interme-
diates in the formation of these two kinds of material.

b1469_Ch-06.indd 307 4/8/2013 12:32:33 PM


b1469 Catalysis by Ceria and Related Materials

308 Z.-X. Li et al.

Figure 6.6 (a) TEM and (b) HRTEM images of 33.8 nm ceria nanoflowers.
(c) HRTEM image of a corner of (b). (d) TEM image of 37.6 nm ceria nanoflowers.
Reprinted with permission from Zhou et al.114 Copyright 2008 Wiley-VCH.

6.2.1.4 Other structures: nanoflowers, hollow particles, and nanoprisms


Ceria with other shapes have also been synthesized as described
in the recent literature. A common shape is the flower. Yan and
co-workers synthesized small ceria nanoflowers in a high-boiling
solvent system (Fig. 6.6).114 The formation mechanism was the ori-
ented attachment of small ceria nanoparticles induced by the rapid
depletion of the nitrate ions as capping agents. The method can also

b1469_Ch-06.indd 308 4/8/2013 12:32:33 PM


b1469 Catalysis by Ceria and Related Materials

2D and 3D Ceria-Based Nanoarchitectures 309

be extended to transition metals to obtain similar oxides with flower


shapes. Lu and co-workers described the hydrothermal synthesis of
ceria flowers composed of 1D nanowires assembled from 0D ceria
nanoparticles.115 The various surfactants were regarded as the main
factor in determining the final morphology. The final product has a
high concentration of Ce3+ and O2− vacancies due to the high sur-
face area/volume ratio of the small nanocrystals. Kempaiah et al.
also described a method of synthesizing ceria microflowers using the
solvothermal method performed in glycol media.116 The flowers
were constructed by the arrangement of ceria nanopetals.
Zhang, Shi, and co-workers also described a method to prepare
a new type of hollow structure assisted by CNTs as templates.117 Ceria
nanocrystals were grown on the CNTs and formed necklace-like
hybrid structures. After calcination, the CNTs were removed to give
ceria nanomaterials with a ring structure.
Hollow ceria nanostructures can be also prepared with the help
of other hard templates. Jian, Du, and co-worker used a silica sphere
as a hard template.118 Ceria nanocrystals were precipitated on the
silica sphere to form a polycrystalline shell. After etching with
NaOH solution to remove the silica, hollow spheres were obtained.
Strandwitz and Stucky also used this method to obtain uniform hol-
low ceria spheres.119
Li and co-workers used carbonaceous polysaccharide micro-
spheres as hard templates.120 The surface of a template was function-
alized with −OH and C=O groups, which can bind with metal cations
through the coordination of electrostatic interactions. After calcina-
tion, the adsorbed metal species formed oxide hollow spheres, while
the carbonaceous spheres oxidized into carbon oxide species.
Recently, Song, Cai, and co-workers used a microwave-assisted
precipitation method to obtain hollow ceria spheres.121 They pro-
posed that the new mechanism was an Ostwald ripening coupled
self-templated self-assembly process. Amorphous solid spheres
formed at the beginning of the reaction stage. In the growth stage,
small nanoparticles grew at the interface between the solid spheres
and the solution. Along with Ostwald ripening, the inner sphere dis-
solved and formed a hollow structure (Fig. 6.7).

b1469_Ch-06.indd 309 4/8/2013 12:32:34 PM


b1469 Catalysis by Ceria and Related Materials

310 Z.-X. Li et al.

Figure 6.7 (a) SEM and (b) TEM images of hollow ceria nanospheres. Reprinted
with permission from Cao et al.121 Copyright 2010 the American Chemical Society.

Yang et al. obtained hollow ceria spheres assembled from octahe-


dra.122 The formation mechanism was explained by Ostwald ripen-
ing during the hydrothermal treatment, which dissolved the inner
cerium species and re-precipitated them outside the particle to form
hollow structures. Sun and co-workers also described a recrystalliza-
tion-based method to produce hollow spheres.123 They employed
PVP as a capping ligand and a water–ethanol mixture as a solvent.
After hydrothermal treatment, hollow spheres composed of ceria
nanopolyhedra were obtained.
Zhang et al. described a synthesis of hollow ceria nanostructures
without using a template or surfactant.124 Single/multiwall hollow
microspheres were obtained by a hydrothermal reaction carried at
230°C for 6–10 hours with a molar ratio of urea to Ce(III) of 3:1 to 6:1.
Sun and co-workers developed a novel template-free method to
prepare hollow ceria nanocubes with peroxyacetic acid as the oxi-
dant.125 The formation mechanism was explained as having two
steps including the oriented attachment of small nanoparticles to
form nanocubes and Ostwald ripening to form the hollow structure
dominated by “solid-solution-solid” mass transportation.
Chen and co-workers described novel prism-like ceria mesocrys-
tals obtained by the hydrothermal method.126 With the correct con-
centration of OH−, the nucleation rate and crystal growth rate are in

b1469_Ch-06.indd 310 4/8/2013 12:32:34 PM


b1469 Catalysis by Ceria and Related Materials

2D and 3D Ceria-Based Nanoarchitectures 311

equilibrium. Small particles were aggregated along specific direc-


tions and formed prism-like mesocrystals.

6.2.2 Assembly of ceria-based nanomaterials


The self-organization of entities in two dimensions (2D) and three
dimensions (3D) is one of the basic processes in nature. Spherical
objects, such as oranges, balls, or particles, with the same diameter
self-organize in compact 2D and 3D ordered structures (face-
centered cubic (fcc) or hexagonal close packed (hcp)). The first
self-assembly of particles having a diameter of a few nanometers
(<10 nm) was discovered in 1995.127,128 The collective properties of
an assembly of nanocrystals are not just a simple sum of the proper-
ties of the isolated particles, because novel properties are gener-
ated by the interaction between particles, which depends on the
shape and nanocrystal structure at the mesoscopic scale. Thus, for
the widespread application of CeO2, most attention has focused on
controlling the size and shape of the primary particles as well as
manipulating their arrangement and aggregation structure. For
instance, thermally stable, nanostructured CeO2 with a high sur-
face area was formed by the self-assembly of individual CeO2 nano-
particles in a liquid crystal phase. It was constructed for an
operational organic-dye-free solar cell and exhibited a photovoltaic
response directly derived from the nanometric structure of the
constituent particles.129 CeO2 with large particles does not have a
photovoltaic response. The efficiency of the material was good and
promises a new type of non-dye-sensitized solar cell after further
optimization of this type of mesostructured material and adequate
doping.

6.2.2.1 Mesoporous structures


Since the first successful synthesis of ordered mesoporous silica,
interest in this research field has expanded all over the world
because of the distinctive properties of mesoporous materials, such
as highly uniform channels, large surface area, narrow pore size

b1469_Ch-06.indd 311 4/8/2013 12:32:34 PM


b1469 Catalysis by Ceria and Related Materials

312 Z.-X. Li et al.

Figure 6.8 TEM images of mesoporous CeO2. (a) Uncalcined ceria–surfactant


composite. (b) Mesoporous ceria after calcination at 673 K in air. (c) Mesoporous
ceria calcined in air at 873 K. Reprinted with permission from Lyons et al.130
Copyright 2002 the Royal Chemical Society.

distribution, tunable pore sizes over a wide range, and so on.


Mesostructured ceria has many excellent properties.
Lyons et al. carefully controlled the reaction conditions and
calcination regimes to prevent complete pore collapse of an as-
synthesized inorganic–organic mesoscopically ordered template
material to allow the formation of ordered mesoporous ceria, using
neutral surfactants (Fig. 6.8). Cerium acetate was chosen as the
inorganic framework precursor and careful thermal processing of
the matrix allowed the subsequent densification of the inorganic
component and removal of the organic component so that high-
quality ordered and truly crystalline mesoporous ceria formed. This
material maintained its high surface area after calcination up to
temperatures of 873 K. The preparation of high-surface-area ceria
with a stable uniform array of pores is significant and may allow the
development of novel catalytic applications for ceria.130 Yu et al. syn-
thesized mesoporous ceria and ceria–zirconia solid solutions with a
high surface area without thermal post-treatment. This structure was
formed by the agglomeration of monodispersed nanoparticles
under high-intensity ultrasound irradiation.131 Si et al. prepared
mesoporous Ce0.2Z0.8O2 solid solutions using a trivalent cerium salt
as the starting material via a hydrothermal method in the presence
of urea without any template. The mesostructure with the highest

b1469_Ch-06.indd 312 4/8/2013 12:32:34 PM


b1469 Catalysis by Ceria and Related Materials

2D and 3D Ceria-Based Nanoarchitectures 313

surface area and largest pore volume had the highest oxygen stor-
age capacity of 233 mmol CO g−1 at 773 K and the lowest 50% con-
version temperature of 641 K.132
Yuan et al. produced mesoporous Ce0.2Zr0.8O2 solid solutions via the
non-template hydrothermal route.133 The mesoporous structure was
fabricated during the nanocrystal growth process and exhibits short-
range order but long-range disorder, which is a result of the natural
agglomeration of uniform nanoparticles after hydrothermal treat-
ment. The as-prepared powders had high surface areas (232–281 m2.g−
1
) and narrow pore size distributions (3.5–4.0 nm). The mesostructured
material with the highest surface area and the largest pore volume had
the highest oxygen storage capacity and the lowest 50% conversion
temperature for CO oxidation. The researchers further developed a
facile method via sol-gel and an evaporation-induced self-assembly
process to synthesize highly ordered 2D hexagonal mesoporous Ce1−
xZrxO2 solid solutions with crystalline walls (Fig. 6.9). The Ce/Zr ratio
can be tuned over a wide range. These novel structured ceria–zirconia
solid solutions are good candidates for application in catalysis due to

Figure 6.9 TEM images of mesoporous Ce1-xZrxO2 (x = 0.5) recorded along the
(a) [001] and (b) [110] orientations. The inset in (a) is the corresponding FFT
(fast Fourier transform) diffraction image, and the one in (b) is the corresponding
SAED pattern. Reprinted with permission from Yuan et al.133 Copyright 2007 the
American Chemical Society.

b1469_Ch-06.indd 313 4/8/2013 12:32:34 PM


b1469 Catalysis by Ceria and Related Materials

314 Z.-X. Li et al.

Figure 6.10 Morphology and structure of CeO2 nanoparticle spherical aggregates


prepared with ILs. (a) SEM image. (b) TEM image. (c) HRTEM image. (d) Schematic.
Reprinted with permission from Li et al.134 Copyright 2008 the American Chemical
Society.

their high surface area and uniform channels. Very recently, Li et al.
made a breakthrough in the self-assembly of CeO2 spherical aggregates
directed by ionic liquids (ILs) (Fig. 6.10),134 which act as both templates
and cosolvent agents in fabricating the ceria spherical aggregates. Li
et al. described a hierarchically macroporous- and mesoporous-
structured cerium–tin mixed oxide with nanocrystalline walls, which
was synthesized through a simple one-step sol-gel process with a block
copolymer as a single template (Fig. 6.11).135 Various kinds of

b1469_Ch-06.indd 314 4/8/2013 12:32:34 PM


b1469 Catalysis by Ceria and Related Materials

2D and 3D Ceria-Based Nanoarchitectures 315

Figure 6.11 Hierarchically macrocellular and mesoporous Ce0.9 Sn0.1O2.


(a) Representative SEM image. (b), (c) Typical TEM images; the inset is the cor-
responding SAED pattern. (d) High-magnification TEM images. Reprinted with
permission from Li et al.135 Copyright 2009 Wiley-VCH.

chemicals, especially those having both hydrophilic and hydrophobic


functional groups, have been mainly used as templates in the assembly
of ceria mesostructures.

6.2.2.2 Other structures


Nanoparticles or colloidal crystals with a uniform shape and size
can be synthesized and used as building blocks for assembled

b1469_Ch-06.indd 315 4/8/2013 12:32:34 PM


b1469 Catalysis by Ceria and Related Materials

316 Z.-X. Li et al.

nanostructures. By employing this strategy, 2D or 3D structures are


easily obtained. Corma et al. prepared a high-surface-area, thermally
stable, nanostructured CeO2 by the self-assembly of individual 5-nm
spherical CeO2 nanocrystals in a liquid crystal phase using a triblock
copolymer as template. The obtained structure had well-defined
hexagonal symmetry.129 Chane-Ching et al. developed a general syn-
thetic method for the synthesis of nanostructured large-surface-area
materials through the self-assembly of functionalized nanoparti-
cles.136 In their work, ceria nanoparticles were functionalized using
bifunctional amino acid species to provide suitable interactions with
the (CH2CH2O) groups of the surfactant. Based upon the coopera-
tive self-assembly of colloidal nanoparticles, hexagonal arrays of
CeO2 were obtained and the symmetry of the arrays was preserved
after heating to temperatures over 500°C (Fig. 6.12). Niederberger
and co-workers created highly ordered 3D CeO2 nanostructures
using a block-copolymer-assisted assembly process.137 The aqueous
CeO2 nanocrystal sol was used as the building blocks without the aid
of external agents such as surface functionalization species or bind-
ing agents.
In contrast to this liquid-crystal template-assisted self-assembly
process, Si et al. found that during a PVP-assisted alcohothermal
synthesis of ceria nanocrystals, three types of self-organized mon-
olayer pattern, namely, short chain-like (pseudo-1D aggregated),
pearl necklace-like (1D aggregated), and dendritic (pseudo-2D
aggregated) alignments, were produced on the copper TEM grids.138
This special self-organization behavior was correlated with the deli-
cate balance between the attractive and repulsive forces caused by
the adsorbed hydroxyls, PVP, and alkylammonium cations on the
surfaces of the CeO2 nanocrystals during the irreversible evapora-
tion of the solvent from various colloidal solutions under ambient
conditions. Yan and co-workers also found 3D oriented attachment
phenomena during the formation of CeO2 nanoflowers in hot sur-
factant solutions by the rapid thermolysis of (NH4)2Ce(NO3)6 oleic
acid (OA)/oleylamine (OM).114 Uniform CeO2 nanoflowers with
controlled shape (cubic, four-petalled, and starlike) and tunable
size (10–40 nm) were obtained by adjusting the reaction conditions

b1469_Ch-06.indd 316 4/8/2013 12:32:35 PM


b1469 Catalysis by Ceria and Related Materials

2D and 3D Ceria-Based Nanoarchitectures 317

(a)

(c)
(b)

Figure 6.12 (a) Self-assembly process of surface-modified ceria nanoparticles.


(b) TEM image of the assembled CeO2 nanostructured material and (c) Small-angle
X-ray scattering pattern. Reprinted with permission from Chane-Ching et al.136
Copyright 2005 Wiley-VCH.

including solvent composition, precursor concentration, reaction


temperature, and reaction time. In situ electrical conductance meas-
urements were carried out to investigate the growth mechanism and
two major steps were confirmed, that is, the initial formation of
ceria cluster particles capped with various ligands (e.g., OA, OM,
and NO3−) via hydrolysis of (NH4)2Ce(NO3)6 at temperatures in the
range 140–220°C, and then secondary self-assembly of the primary
particles into nanoflowers above 220°C, which was ascribed to
the significant reduction in surface ligand coverage caused by the
abrupt decomposition of (NH4)2Ce(NO3)6 at the elevated

b1469_Ch-06.indd 317 4/8/2013 12:32:35 PM


b1469 Catalysis by Ceria and Related Materials

318 Z.-X. Li et al.

temperature during a strong redox reaction. As discussed above,


oleic acid-assisted hydrothermal synthesis method is a facile strategy
for the fabrication of monodisperse nanocrystals with narrow size
distributions and controlled shapes, which can serve as ideal build-
ing blocks for a bottom-up approach in forming various 2D or 3D
superstructures. Li and co-workers described a pathway for the mul-
tigram synthesis of ultra-fine monodisperse CeO2 nanocrystals in a
water–ethanol mixed solvent using this method. These CeO2
nanocrystals were spontaneously assembled to form submicrometer
3D colloidal crystals with highly ordered superlattice structures in
one step without further treatment.139
Sayle et al. developed the classical atomistic simulation to
produce combined studies of theoretical and experimental works
(Fig. 6.13). A typical selected system was ceria. Since the pair
potential model based on electrostatic interaction and Buckingham

Figure 6.13 Nano-building blocks. (a) HRTEM of a CeO2 nanocrystal. (b) Three-
dimensional tomogram of a CeO2 nanocrystal generated from computer-aided
tomography of sequentially oriented TEM images. (c) Atomistic model of a CeO2
nanocrystal. (d) Ti-doped CeO2 nanocrystal showing the (amorphous) TiO2 shell
encapsulating the inner (crystalline) CeO2 core rendering it spherical. (e)
Atomistic model of a Ti-doped CeO2 nanocrystal. Reprinted with permission from
Sayle et al.140 Copyright 2008 the American Chemical Society.

b1469_Ch-06.indd 318 4/8/2013 12:32:35 PM


b1469 Catalysis by Ceria and Related Materials

2D and 3D Ceria-Based Nanoarchitectures 319

short-range presentations are often adequate to describe the fluo-


rite structure of ceria, they explored the application of such
models to nanosized particles. A series of works has described the
assembly behavior of nano-building blocks into complex
nanostructures.140
Karakoti et al. suggested a possible strategy for making 1D self-
assembled nanostructures by exploiting natural phenomena, such
as voids formed during the freezing of water.141 During freezing of
a dilute aqueous solution of CeO2 nanocrystals, some nuclei
remain in solution while others are trapped inside micrometer
and nanometer voids formed within the growing ice front. Over
time (2–3 weeks) the particles trapped within the nanometer-wide
voids in the ice combine by an oriented attachment process to
form ultra-long polycrystalline ceria nanorods with a length of ca.
3–3.5 mm and diameter of ca. 30 nm (Fig. 6.14(a)). Constant-
volume molecular dynamics (MD) simulation revealed attractive
forces acting between neighboring amorphous/molten nanoparti-
cles (Fig. 6.14(b)). Accordingly, under MD simulation, the nano-
particles start to move closer to one another and began to
crystallize. Once crystalline, the nanoparticles continue to be
attractive towards one another. By simulating self-assembly dynam-
ically, similar to experiments, the models derived are metastable
and do not thermodynamically represent the lowest-energy struc-
tures. This work has shown that a natural phenomenon, such as
the structural formation of ice, can architecturally sculpt 1D ceria
nanostructures. MD simulations confirm the type of morphology
evolution observed for the self-assembled ceria nanostructures
comprising nanoparticle building blocks.
Kuchibhatla et al. prepared a hierarchically assembled octahe-
dral superstructure of CeO2 from nanoscale building blocks with-
out any template or external driving forces.142,143 They proposed
that the dynamic self-assembly creates an intra-agglomerate reori-
entation to attain a low energy configuration after long-term
aging of the nanoparticles. In order to explain the formation
process, they used MD simulation, HRTEM, and fast Fourier
transforms (FFT). The results proved to be vital in confirming

b1469_Ch-06.indd 319 4/8/2013 12:32:36 PM


b1469 Catalysis by Ceria and Related Materials

320 Z.-X. Li et al.

Figure 6.14 (a) Image generated by selected masking and inverse FFT of a
HRTEM of CeO2 nanorods showing the dislocations (circles) formed during orien-
tation and self-assembly. (b) Structure of a chain/nanorod comprising CeO2
nanocrystals. Top: After 300 ps of MD simulation; the arrows indicate the trajectory
followed by the blue and green nanoparticles as they attach themselves to one
another. Bottom: After 1500 ps showing the attached configuration. Reprinted with
permission from Karakoti et al.141 Copyright 2008 Wiley-VCH.

b1469_Ch-06.indd 320 4/8/2013 12:32:36 PM


b1469 Catalysis by Ceria and Related Materials

2D and 3D Ceria-Based Nanoarchitectures 321

Figure 6.15 Representations of the atom positions of a model ceria nanoparticle


after attachment. (a) Three ceria nanoparticles attached at {100}; the coherent
interfaces are shown more clearly in (b). (c) Schematic detailing the behavior.
(d) Enlarged interfacial region, revealing the considerable transport of ions (high-
lighted by the gray rectangle) to the interfacial region during the simulation to
facilitate a coherent interface. Reprinted with permission from Kuchibhatla et al.142
Copyright 2009 the American Chemical Society.

the formation of superoctahedra as a result of the fractal agglom-


eration of nanoscale octahedral building blocks of ceria. With the
help of FFT, the orientation of the nanoparticles within the
agglomerates and their interfacial defects were analyzed. The nan-
oparticles undergo an intra-agglomerate rotation to align in a
preferred direction while forming the final morphology (Fig. 6.15).

b1469_Ch-06.indd 321 4/8/2013 12:32:37 PM


b1469 Catalysis by Ceria and Related Materials

322 Z.-X. Li et al.

6.3 Properties and Applications


Compared with microsized or bulk materials, CeO2 nanostructures
show significant size-induced property changes in a wide range of
aspects, such as lattice expansion, blue shift in ultraviolet absorption
spectra, Raman-allowed modes shifting and broadening, and pres-
sure-induced phase transformation. The lattice relaxation can be
observed to be size-induced in ceria nanoparticles with average sizes
4–60 nm.144 Diffraction line broadening is attributed to the size and
strain of the ceria nanoparticles. For fine crystals (<5 nm), the size
contributes to the line broadening, while the strain effect is signifi-
cant for larger crystals. The formation of oxygen vacancies and the
associated Ce3+ result in the shift in lattice parameters with crystal
size. When the crystal size increased from 4 to 60 nm, the concentra-
tion of oxygen vacancies changed (by two orders of magnitude).
Wang et al. described pressure-induced effects in nanoceria, such as
a phase transition and size changes.145,146 The phase transition was
observed by in situ synchrotron X-ray diffraction when studying size-
induced compressional effects in nanocrystalline CeO2 up to a pres-
sure of 38 GPa. The results indicate that at a critical pressure of 20
GPa, a significant weakening of the size-induced effect occurs.
Tsunekawa et al. described electron diffraction studies on anomalous
lattice expansions in monodisperse CeO2−x nanoparticles.147 X-ray
photoelectron spectroscopy (XPS) results suggest that a valence
reduction of Ce ions leads to a decrease of the electrostatic force and
therefore is the origin of the expansion. Zhang et al. studied the size-
dependent ultraviolet (UV) absorption by ceria nanocrystals. The
blue-shifting in the UV absorption spectra was observed for CeO2−δ
nanocrystals for both direct and indirect bandgap energies in the
range of 4–7 nm due to the quantum size effect. The bandgap ener-
gies were almost unchanged for ceria nanocrystals with sizes less than
4 nm probably because the blue-shifting due to the quantum size
effect was counteracted by red-shifting due to the dielectric confine-
ment effect.148 Tsunekawa et al. also investigated the UV absorption
spectra of monodisperse CeO2−x (0 < x < 0.5) nanocrystals. The
nanocrystals showed blue shifts with an absorption edge about
310–400 nm, which could be related to the particle sizes. They

b1469_Ch-06.indd 322 4/8/2013 12:32:38 PM


b1469 Catalysis by Ceria and Related Materials

2D and 3D Ceria-Based Nanoarchitectures 323

proposed that the blue shift is due to the valence change of the Ce
ions.149 Spanier et al. reported on Raman spectra studies of CeO2−y
nanoparticles. As the particle size becomes smaller, the peak position
of the strong, triply degenerate, first-order Raman line in CeO2−y at
464 cm−1 is shifted to progressively lower energies and the peak gets
progressively broader and asymmetric on the low-energy side.150 The
size-dependent properties of ceria nanoparticles are explained by
the combined effect of strain and phonon confinement effects. The
lattice constants increased with a decrease in particle size, and there-
fore, the Raman peaks shift to low energies. The line-width change
was explained by inhomogeneous strain broadening with a small
dispersion in particle size and by phonon confinement.

6.3.1 Catalysis
In most catalytic reactions, CeO2 acts as an oxygen buffer by storing
and releasing O2 due to the Ce4+/Ce3+ redox couple, which is a
result of the intrinsic defect chemistry of ceria (Ce3+ and O vacancy
concentrations). In simulation studies, the role of CeO2 in the oxi-
dation of carbon monoxide (CO) has been investigated and this
reaction can be described, following Guo et al.,118 and the oxidation
of CO to CO2 can be described, following Sayle et al.151 as:

CO(g) + 2CeCeX + OOX → CO2(g) + VO″ + 2CeCe′

where oxygen is extracted from an exposed surface of the CeO2


facilitated by the reduction of two Ce4+ ions to two Ce3+ ions.
Simulations reveal that oxygen vacancy formation, together with
Ce4+ reduction to Ce3+, is energetically more favorable on the sur-
face of the CeO2 compared with the bulk. Moreover, the facile for-
mation of oxygen vacancies on the surfaces of CeO2 significantly
promotes the oxidation of CO.151,152 In some works, the Ce4+/Ce3+
ratio has been reported as varying with particle size or in correlation
with other species.153 A pre-designed synthesis with a precisely
adjusted Ce4+/Ce3+ ratio, the defect content, is required for special
applications in catalysis.

b1469_Ch-06.indd 323 4/8/2013 12:32:38 PM


b1469 Catalysis by Ceria and Related Materials

324 Z.-X. Li et al.

A number of reactions could be catalyzed by ceria and ceria-


based catalysts (ceria–noble metal, ceria–transition metal oxides),
such as the water gas shift (WGS) reaction, CO selective oxidation
by oxygen, the catalytic combustion of volatile organic compounds,
selective dehydrogenation processes, as well as CO oxidation in
automobile waste gas. Therefore, investigations of catalysis by ceria
nanomaterials are primarily about the catalytic oxidation of CO in
various gas mixtures and conditions. Fuel cells are an attractive new
power-generation technology to convert chemical energy directly
into electricity with high efficiency and low pollution. A solid oxide
fuel cell (SOFC) operates with three processes: the reduction of
molecular O2 at the cathode, the diffusion of O2− anions through an
oxygen ionic electrolyte, and the oxidation of fuel at the anode.
Ceria has proved to be a useful component of anode materials when
the self-cleaning of hydrocarbon fuels under appropriate operating
conditions is required.154,155 Cerium oxide and lanthanide oxides are
used to adsorb and remove H2S from a fuel gas stream to protect the
anodes from sulfurization.156 Ceria or doped ceria nanomaterials
can be used as a buffer layer between the electrodes and electrolytes
to improve ionic conductivity.157 Nanostructured electrodes in fuel
cells provide advantageous catalytic properties due to the enhanced
surface vacancy concentration and increased ionic and electronic
conductivities. Though the stability of the nanostructured compo-
nents under elevated operating temperatures raises issues of con-
cern, extended stability has been shown over hundreds of cycles in
fuel cell systems.158 As well as their use in SOFCs, ceria nanomateri-
als have more important applications in fuel pretreatment by cataly-
sis in WGS and selective CO oxidations.
Ceria–noble metal (such as Ru, Rh, and Pd) catalysts are com-
posed of noble metal species such as nanoparticles and clusters dis-
persed on the ceria supports. The catalysts show typical strong
metal/support interactions (SMSIs).159 For example, the catalysts
exhibit a number of features for SMSI effects including: (i) reducible
supports; (ii) “high-temperature” reduction treatments; (iii) heavily
disturbed chemical properties and significant changes in catalytic
behavior of the dispersed metal phase; (iv) reversibility in recovering

b1469_Ch-06.indd 324 4/8/2013 12:32:38 PM


b1469 Catalysis by Ceria and Related Materials

2D and 3D Ceria-Based Nanoarchitectures 325

the conventional behavior of the supported metal phase. The reduc-


ibility of ceria nanoparticles is greatly enhanced by the noble metal
species and the catalytic activity of the noble metals is enhanced by
ceria nanoparticles.
The water gas shift reaction (CO+H2O ↔ H2+CO2) is one of the
critical processes for providing hydrogen for industrial applications.
Ceria–noble metal composite catalysts have proved to be a competi-
tive alternative for traditional catalysts in low-temperature water gas
shift (LTS) reactions. Flytzani-Stephanopoulos and co-workers made
a series of studies on nanostructured ceria–gold composite catalysts
for LTS.160–162 The composite was made of ceria or La-doped ceria
NPs and gold NPs or ionic species. Three different preparation
methods were used: deposition-precipitation,163,164 co-precipitation,
and urea-assisted gelation.160 They found that with a gold loading of
5–8 wt%, the composite is a very good catalyst for LTS. Temperature-
programmed reduction (TPR) results showed that the addition of
gold significantly increases the reducibility of the ceria surface oxy-
gen, independent of the different preparation routes. The CO con-
version profile with temperature showed that the conversion rates
were affected by the loading amount of Au and the crystal size of the
ceria. With increased gold loading and smaller ceria size, WGS activ-
ity increased while gold particle size did not have a significant effect
on catalysis activity.160 However, without the gold species, the activity
of ceria NPs is much lower. The WGS light-off temperature of gold
or copper-modified ceria samples is below 120°C, while ceria itself is
inactive below 300°C.161 XPS tests suggested that in the nanostruc-
tured ceria-gold catalysts, most of the gold species are metallic; how-
ever, there are also ionic species. Fu et al. tested the significance of
ionic species for catalysis by leaching gold metal NPs away from the
ceria–gold NP composite catalyst. The results are striking in that the
catalytic activity for the WGS with or without gold NPs is the same,
given that ionic gold species exist. Therefore, it is suggested that
non-metallic Au species instead of NPs play the major role in the
WGS process catalyzed by Au–ceria catalysts.162,165 Since only non-
metallic gold species have a strong effect on the reaction, a much
lower gold content could be used in this catalyst for economic

b1469_Ch-06.indd 325 4/8/2013 12:32:38 PM


b1469 Catalysis by Ceria and Related Materials

326 Z.-X. Li et al.

(a)

(b)

Figure 6.16 Composition and activity for the WGS of low content gold–ceria catalysts.
(a) Amount of gold before and after leaching at different calcination temperatures.
(b) Steady-state WGS reaction rates. Solid triangles: 4.7Au–10% La doped ceria (dep-
osition-precipitation); open triangles: leached 0.44Au – 10%La doped ceria (deposi-
tion-precipitation, NaCN). Other reference catalysts are also shown: solid circles:
0.62Au–TiO2 (deposition-precipitation), solid squares: 2.02Au–Fe2O3 (co-precipita-
tion), open squares: leached 0.73Au–Fe2O3 (co-precipitation, NaCN). Reprinted with
permission from Flytzani-Stephanopoulos.163 Copyright 2004 Elsevier.

reasons. A low gold content catalyst is free of metallic gold NPs while
exhibiting high catalytic activity (Fig. 6.16).165
The known WGS catalysts based on ceria or copper oxide both
suffer greatly from deactivation with time on-stream and in

b1469_Ch-06.indd 326 4/8/2013 12:32:38 PM


b1469 Catalysis by Ceria and Related Materials

2D and 3D Ceria-Based Nanoarchitectures 327

shutdown/restart operations. The mechanism for this deactivation


was ascribed to the blocking of active sites by carbonates or formates
formed during the WGS reaction or by over-reduction of ceria.
Flytzani-Stephanopoulos et al. studied the deactivation of ceria–gold
or ceria–platinum composite catalysts in a practical fuel cell.166 The
formation of cerium(III) hydroxycarbonate in the ceria-gold cata-
lysts was identified during the shutdown process. Therefore, the
addition of a small amount of gaseous oxygen to the reaction gas
mixture was proposed as a solution. Both Au–ceria and Pt–ceria
catalysts showed remarkable stability in the WGS reaction in a full
fuel-gas mixture at various temperatures and in cyclic shutdown/
start-up operations. The selective oxidation or preferential oxida-
tion (PROX) of CO in a hydrogen-rich stream is another important
goal for ceria-based catalysts. The gas mixture from steam reforming
or partial oxidation of alcohols and hydrocarbons followed by
the WGS reaction contains mainly H2, CO2, and a small portion of
CO, H2O, and N2. When this type of gaseous stream is used as input
for hydrogen fuel cells, the CO has to be removed to avoid poison-
ing the anode electrocatalysts. Ceria-based nanomaterials, such as
ceria–gold and ceria–copper oxide catalysts, exhibit suitable cata-
lytic activities and selectivity for the CO PROX process.
Corma and co-workers reported that the activity of gold on
nanocrystalline CeO2 for CO oxidation increased by two orders of
magnitude with respect to a conventionally precipitated CeO2 sup-
port. The characteristics of the cerium oxide surface are extremely
important in determining whether a CeO2-supported gold catalyst is
active or not for CO oxidation.31 Catalysts of nanoparticulated gold
supported by nanocrystalline or mesostructured nanocrystalline ceria
when used for the selective aerobic oxidation of aliphatic and aro-
matic aldehydes are much better than gold catalysts supported by
precipitated ceria.167 The researchers also used the Au/
mesostructured CeO2 catalyst to oxidize 4-iso-propylbenzaldehyde to
cumic acid (4-isopropylbenzoic acid), which is an intermediate in the
manufacture of nateglinide. The results show that Au/mesostruc-
tured CeO2 is able to perform the reaction with air at atmospheric
pressure with high conversion and selectivity. Nanoparticulated Au on
nanocrystalline or on mesostructured nanocrystalline CeO2 supports

b1469_Ch-06.indd 327 4/8/2013 12:32:38 PM


b1469 Catalysis by Ceria and Related Materials

328 Z.-X. Li et al.

is an alternative to actual metal soluble or heterogeneous catalysts for


the selective oxidation of aldehydes to carboxylic acids by air.167 And
they carried out time resolved spectroscopy studies on the working
catalysts of gold nanocluster/nanocrystalline ceria.168 CO-TPR, XPS,
in situ Raman, and IR spectra proved that nanocrystalline CeO2 sup-
plies reactive oxygen in the form of surface superoxide species and
peroxide species at one-electron defect sites to the supported active
species of gold for the oxidation of CO. Reactive oxygen species are
not formed on conventionally prepared CeO2 and the presence of the
gold enhances the formation of the reactive oxygen species on
nanocrystalline ceria. In addition, Corma and co-workers reported
that nanocrystalline CeO2 and Y2O3 could stabilize the Au(III) species
on the surface and that the catalytic activity of CO oxidation is propor-
tional to the concentration of these Au(III) surface species.169
Nanocrystalline CuO–CeO2 catalysts show nearly ideal selectivity
and promising activity for the removal of CO from reformed fuels by
selective oxidation. Hočevar and co-workers studied the kinetics of
CO selective oxidation in excess hydrogen over a nanostructured
Cu0.1Ce0.9O2−y catalyst prepared by a sol-gel method.170 A steady-state
Mars and van Krevelen kinetic model was proposed for the reaction
and the kinetic parameters for the reaction were found to be as fol-
lows: the apparent activation energy for CO oxidation, 57.2 kJ.mol−1,
and for catalyst reoxidation, 60.2 kJ.mol−1. The ultra-fine nanocrys-
talline CuO–CeO2 catalysts were made by a urea-nitrate combustion
method.171 CuO-CeO2 catalysts are inactive for H2 oxidation at tem-
peratures up to ca. 120°C and CO does not influence the rate of
hydrogen oxidation. The addition of small amount of H2O has an
adverse influence on CO oxidation. The catalysts also showed stable
activity in long-term experiments with realistic feeds. Ratnasamy
et al. described CuO-CeO2-ZrO2 nanocrystalline catalysts made by a
co-precipitation method for CO selective oxidation.172 The composi-
tion of the support markedly influenced PROX activity. Both CuO–
CeO2 and CuO–CeO2–ZrO2 catalysts exhibited higher activity and
selectivity in CO oxidation than CuO–ZrO2.
Ceria- or yttria-supported Au catalysts are also active and
extremely selective for the homocoupling of arylboronic acids, and

b1469_Ch-06.indd 328 4/8/2013 12:32:38 PM


b1469 Catalysis by Ceria and Related Materials

2D and 3D Ceria-Based Nanoarchitectures 329

the activity is directly correlated with Au(III).169 Other oxidation or


reduction related reactions with ceria-based catalysts were also
explored. For example, Murugan and Ramaswamy described the
oxidative dehydrogenation of ethylbenzene on nanocrystalline ceria
using N2O as the oxidant.173 Concepcion et al. described the chem-
oselective hydrogenation of crotonaldehyde catalyzed by Pt on meso-
structured CeO2 NPs embedded within layers of an SiO2 binder.174
Cationic and metallic gold are present in gold catalysts prepared
with nanocrystalline CeO2 and the catalytic active sites incorporate
cationic gold. Nanocrystalline CeO2 stabilizes O2 as superoxide and
peroxide species, whereas precipitated CeO2 tends to stabilize O2
and molecular O2. Moreover, nanocrystalline CeO2 supplies reactive
oxygen in the form of surface n1 superoxide species and peroxide
adspecies at one-electron defect sites to the supported active species
of gold for the oxidation of CO.175
Since the advent of three way catalysts (TWCs) for vehicle waste
gas treatment in the 1980s the demand for removing pollutants has
increased. TWCs remove pollutants by simultaneously converting
nitrogen oxides, carbon monoxide, as well as unburnt hydrocarbons
into harmless H2O, CO2, and N2. TWCs perform most efficiently
when the air-to-fuel ratios are equal to the stoichiometric conditions,
and deviations from the ideal ratios severely decrease the efficiency
of TWCs. The addition of CeO2 could limit this disadvantage by act-
ing as an oxygen buffer, which stores and releases O2 using the Ce4+/
Ce3+ redox couple.22 The defective structure of nanocrystalline ceria-
based catalysts has been proved to have a strong effect on the OSC.
Mamontov et al. described neutron diffraction studies of the atomic
structure of a nanocrystalline powder of ceria and ceria–zirconia
solid solution.176 They found that the concentration of vacancy-
interstitial oxygen defects has a direct correlation with the OSC.
This effect is stronger than the correlation of surface area with OSC.
Zirconia reduces ceria and preserves oxygen defects to retard the
degradation of ceria–zirconia in OSC.
Yan and co-workers prepared Ce1−xZrxO2 (x = 0−0.8) nanopar-
ticulate powders by a mild urea hydrolysis-based hydrothermal
method through homogeneous nucleation at 413 K followed by

b1469_Ch-06.indd 329 4/8/2013 12:32:38 PM


b1469 Catalysis by Ceria and Related Materials

330 Z.-X. Li et al.

calcination at 773 or 1173 K. They observed that the Ce1−xZrxO2 (x =


0–0.8) nanoparticulate powders maintained a rather high strain
level in the crystal lattice according to powder X-ray diffraction data
and HRTEM images. The lattice strains of these Ce1−xZrxO2 catalysts
have a linear relation with their oxygen storage capacity. This linear
correlation was explained on the basis of a qualitative structural and
defect analysis (Fig. 6.17).177
In addition to the effect of the size of the ceria nanomaterials on
catalytic activity, the exposed specific crystal facets of ceria nanocrys-
tals also play a role.95,178,179 Oxygen storage takes place both at the
surface and in the bulk for nanorods and nanocubes, but it is
restricted to the surface for nanopolyhedra as its bulk counterpart,
due to the exposure of more reactive (100)/(110) planes for the
former two samples. This shape-dependent OSC behavior strongly
suggests that high OSC ceria materials for TWCs can be designed by
tuning their shapes with specific surface crystallographic facets,
such as by increasing the fraction of the more reactive {100} and
{110} planes, together with decreasing the fraction of the less
reactive {111} planes in nanocrystalline catalysts.95 A gold catalyst
supported by ceria for the water gas shift reaction showed a

(a) (b)

(c) (d)

Figure 6.17 HRTEM images of Ce1-xZrxO2 samples: (a) x = 0, calcined at 773 K;


(b) x = 0.5, calcined at 773 K; (c) x = 0, calcined at 1173 K; (d) x = 0.5, calcined at
1173 K. Right: CO-OSC values of Ce1−xZrxO2 (x = 0-0.8) catalysts as a function of
microstrain in the crystal lattice. Reprinted with permission from Si et al.177
Copyright 2004 the American Chemical Society.

b1469_Ch-06.indd 330 4/8/2013 12:32:38 PM


b1469 Catalysis by Ceria and Related Materials

2D and 3D Ceria-Based Nanoarchitectures 331

(a) (b)

Figure 6.18 (a) Two-step synthesis of Au on CeO2 nanorods, nanocubes, and


nanopolyhedra. (b) WGS reaction “light-off” profiles for 1% Au on CeO2 nanorods
(•), nanocubes (ƒ), and nanopolyhedra (S). Reprinted with permission from Si
and Flytzani-Stephanopoulos.179 Copyright 2008, Wiley-VCH.

dependence on the exposed facets of the nanoceria relating to the


shape of the ceria nanocrystals (Fig. 6.18).179 Rodriguez and co-
workers investigated the WGS reaction with Cu and Au NPs sup-
ported on CeO2(111) and ZnO(000-1) surfaces experimentally and
theoretically. Au/CeO2(111) was an excellent catalyst with activity
similar to that of Cu/CeO2(111). The most difficult step in the WGS
reaction is the dissociation of H2O, which can be realized when an
O vacancy exists on the ceria (111) surface. The metal NPs facilitate
the partial reduction of ceria by CO and help to form the O vacan-
cies, illustrating a cooperative effect.180
Corma and co-workers described chemoselectivity exhibited by
gold NPs supported by nanocrystalline ceria for the aerobic oxida-
tion of allylic alcohols.181 Pd NPs and Au/Pd core/shell NPs were
also examined and the Au NPs were more active and chemoselective
when both solvent free and in organic media. The results show that
the gold/ceria system is a general and reusable catalyst for the oxi-
dation of allylic alcohols without a solvent or in organic media.
Catalytic activity in toluene correlates linearly with the total number
of external gold atoms and with the surface coverage of the support-
ing ceria nanostructures.182
Pt/CeO2 hetero-nanocomposites were prepared from Pt/CeO2@
SiO2 obtained by a microemulsion-mediated method.183 Facilitated
by the earlier calcination under the protection of a silica shell, the

b1469_Ch-06.indd 331 4/8/2013 12:32:39 PM


b1469 Catalysis by Ceria and Related Materials

332 Z.-X. Li et al.

as-formed Pt/CeO2 hetero-nanocomposites exhibit good thermal


stability, which can preserve their pristine properties after subse-
quent calcination at temperatures as high as 450 °C. The thermally
stable Pt/CeO2 hetero-nanocomposites have a small particle size,
low aggregation, and maximized Pt/CeO2 interfaces, and thus
exhibit high catalytic activity for CO oxidation. This approach is
simple albeit efficient and can be potentially extended to the synthe-
sis of other composites of an oxide and noble metal (Fig. 6.19).

6.3.2 Sensors
The fundamental sensing mechanism for metal oxide-based gas sen-
sors relies on a change in electrical conductivity due to the interaction

Figure 6.19 (a) Synthesis protocol for Pt/CeO2@SiO2 nanospheres and Pt/CeO2
hetero-nanocomposites. (b) CO conversion vs. reaction temperature over Pt/CeO2
hetero-nanocomposite catalysts and two mixtures of CeO2 and Pt. Reprinted with
permission from Zhou et al.183 Copyright 2010 the American Chemical Society.

b1469_Ch-06.indd 332 4/8/2013 12:32:39 PM


b1469 Catalysis by Ceria and Related Materials

2D and 3D Ceria-Based Nanoarchitectures 333

between surface complexes of O−, O2−, H+, and OH− reactive chemical
species and the gas molecules to be detected. The redox catalytic activ-
ity and the semiconductivity of ceria allow it to be used as a gas sensor
for reductive and oxidative gases such as CO, NO2, O2, and alcohols
using resistivity or cataluminescence measurements. The effects of the
microstructure, namely, the ratio of surface area to volume, grain size,
and pore size on gas-sensing performance is of significant importance,
since reactions at the grain boundaries and a significant depletion of
carriers in anisotropic nanosystems can strongly modify the redox and
transport properties, leading ultimately to the development of quan-
tum-confined structures.12–20 Hence, a ceria nanostructure could also
be used in a gas sensor. In order to increase the sensitivity of ceria for
gas sensing, noble metals or metal oxides are usually applied to acti-
vate the ceria catalytic process.
Liao and co-workers successfully fabricated a gas sensor using a
single ceria nanowire (NW)33 as the sensing unit. The gas sensitivity
of this nanodevice can be improved by depositing Pt nanoparticles
on the ceria nanowire. The nanodevice exhibited high selectivity in
detecting CO over other reductive gases, such as H2, H2S, ethanol,
and gasoline. The researchers thought that the strong adsorption of
the Pt nanoparticles on the ceria nanowire and the charge transfer
between the Pt nanoparticles and the ceria nanowire were the most
important contributions to the high gas sensitivity of this Pt/CeO2
nanowire sensor. Based on the excellent chemical stability of ceria
and its high gas-sensing behavior, Pt/CeO2 nanowire sensors are
expected to have promising applications in various complex envi-
ronments for the fast detection of poisonous gases (Fig. 6.20).33
Fu et al. described fast humidity sensors based on ceria nano-
wires prepared by a hydrothermal method. Sensitivity increased
gradually as the humidity increased and resistance decreased expo-
nentially with increasing humidity. The humidity sensing mecha-
nism was ion-type conductivity, and a model based on the morphology
and surface energy of the nanowires was used to explain these
results (Fig. 6.21).184
Barreca et al. prepared columnar ceria nanostructures on Si(100)
or Al2O3 substrates by a catalyst-free CVD process at 623–723 K using

b1469_Ch-06.indd 333 4/8/2013 12:32:39 PM


b1469 Catalysis by Ceria and Related Materials

334 Z.-X. Li et al.

Figure 6.20 (a) SEM image of a single CeO2 NW gas sensor, inset is an optical
image of the NW sensor chip. I-V curves of a single CeO2 NW and a single Pt NC/
CeO2 NW (b) before and (c) after annealing at 400°C for 1 h. (d) Gas sensibility
response curves of a single CeO2 NW and a single Pt NCs/CeO2 NW, after being
annealed, for different CO concentrations. Reprinted with permission from Liao
et al.33 Copyright 2008 the American Chemical Society.

Figure 6.21 (a) Characteristics of a humidity sensor at different humidities at V = 1


V (curve a: 33% RH; curve b: 54% RH; curve c: 75% RH; curve d: 85% RH; and curve
e: 97% RH). (b) Response and recovery of humidity sensors of CeO2 pore. (c) Scheme
of the absorption and desorption of water molecules in the nanowire and the pore.
Reprinted with permission from Fu et al.184 Copyright 2008 IOP Publishing Ltd.

b1469_Ch-06.indd 334 4/8/2013 12:32:39 PM


b1469 Catalysis by Ceria and Related Materials

2D and 3D Ceria-Based Nanoarchitectures 335

Ce(hfa)3×diglyme (hfa = 1,1,1,5,5,5-hexafluoro-2,4-pentanedione;


diglyme = bis(2-methoxyethyl)ether) as the cerium molecular
source.185,186 Compared with continuous ceria thin films, the ceria
columnar nanostructures showed higher sensitivity in the detection
of gaseous ethanol and NO2.
Glucose biosensors are important for monitoring blood glucose
for treating both diabetics and non-diabetics and could be fabri-
cated through the immobilization of glucose oxidase (GOx). Ceria
is known to have a wide bandgap of 3.4 eV, high isoelectric point of
9.0, which is an advantage for the immobilization of a low isoelec-
tric-point enzyme such as glucose oxidase (isoelectric point 4.2) via
electrostatic interactions with good retention of biological activity
for protein binding. Ansari et al. fabricated a glucose biosensor
based on a GOx immobilized nanostructured CeO2 film deposited on
an Au electrode using a sol-gel method.187 This nanostructured GOx/
CeO2/Au bioelectrode exhibited a sensitivity of 0.00287 µA.mg.dL−1.
cm2. Saha et al. also described a glucose sensor based on a nanopo-
rous CeO2 thin film deposited onto a Pt-coated glass plate with glu-
cose oxidase immobilized using pulsed laser deposition (PLD).188
Guo and co-workers described a nanocomposite membrane of
nanosized shuttle-shaped ceria,189 single-walled carbon nanotubes
(SWNTs), and ILs 1-butyl-3-methylimidazolium hexafluorophos-
phate (BMIMPF6) on a glassy carbon electrode (GCE) for the elec-
trochemical sensing of the immobilization and hybridization of
DNA.190 The electron transfer resistance (Ret) of the electrode sur-
face increased after the immobilization of probe ssDNA on the
CeO2–SWNTs–BMIMPF6 membrane, which improved the sensitivity
in detecting the target DNA.
Cataluminescence (CTL) is the chemiluminescence produced
during catalytic oxidation reactions. Ye et al. studied CTL and the cata-
lytic oxidation reactions of ethanol on nanosized Ce1−xZrxO2 materials.
The ceria-rich solid solutions (x = 0.05–0.25) showed high CTL activ-
ity at 220°C and were considered to be efficient low-temperature CTL
sensors for ethanol. The CTL intensity of ethanol oxidation was
closely related to the steady-state activity of the Ce1−xZrxO2 catalysts in
the oxidative dehydrogenation of ethanol.191 Xuan et al. studied the
CTL of CS2 on the surface of ceria nanorods, nanocubes, and

b1469_Ch-06.indd 335 4/8/2013 12:32:40 PM


b1469 Catalysis by Ceria and Related Materials

336 Z.-X. Li et al.

nanospheres. The chemiluminescence was used as a sensitive gas


sensor for the determination of CS2 and a high selectivity over a
number of organic gases was demonstrated.192

6.3.3 Biomedical applications


Catalytic applications are now the most important use for ceria-based
nanomaterials. However, the intrinsic Ce4+/Ce3+ change properties of
ceria suggest that it can also fully use its capabilities in biomedicine.
Ceria has an astonishing pharmacological potential due to its antioxi-
dant properties, non-toxicity, and excellent biocompatibility, deriving
from a fraction of the Ce3+ ions present in ceria. These defects, com-
pensated by oxygen vacancies, are enriched at the surface and there-
fore in nanosized particles. Reactions involving redox cycles between
the Ce3+ and Ce4+ oxidation states allow ceria nanoparticles to react
catalytically with superoxide and hydrogen peroxide, mimicking the
behavior of two key antioxidant enzymes, superoxide dismutase and
catalase, potentially abating all noxious intracellular reactive oxygen
species (ROS) by a self-regenerating mechanism. Therefore, ceria
nanoparticles, apparently tolerated well by organisms, might fight
chronic inflammation and the pathologies associated with oxidative
stress, which include cancer and neurodegeneration.193
The main ROS are superoxide (O2−), hydrogen peroxide
(H2O2), and the hydroxyl radical (·OH). The Ce3+/Ce4+ transfer in
ceria nanoparticles, similar to redox enzymes, catalyzes reversible
redox reactions in cells and tissues. Pioneering research has revealed
that ceria nanoparticles could be effective in inhibiting the progres-
sion of reactive oxygen intermediates, which induce cell death,194
and so could be useful in therapy for diseases that cause blindness.
Nanoceria was also found to be effective for spinal-cord repair, treat-
ing other diseases of the central nervous system,195 and for protect-
ing normal human cells from irradiation.196 The potential of
vacancy-engineered ceria in disease therapy and anti-aging are suf-
ficiently fascinating to push us towards more research in this field.
Tarnuzzer et al. described vacancy-engineered ceria nanostruc-
tures for the protection of normal human cells from irradiation.197

b1469_Ch-06.indd 336 4/8/2013 12:32:40 PM


b1469 Catalysis by Ceria and Related Materials

2D and 3D Ceria-Based Nanoarchitectures 337

The ceria nanoparticles provided almost 99% protection from


radiation-induced cell death for normal cells, whereas the same con-
centration gave almost no protection to tumor cells. Photoreceptor
cells were exposed to light and had a high rate of oxygen metabo-
lism. These cells are exposed to continuous elevated levels of toxic
reactive oxygen intermediates. Although the causes of retinal dis-
ease resulting in the loss of vision are complex, oxygen radicals,
which damage the sensitive cells in the retina, are thought to play a
central role. Chen et al. showed that ceria nanoparticles could be
effective in inhibiting the progression of reactive oxygen intermedi-
ates responsible for induced cell death (Fig. 6.22). The vacancy
engineered mixed-valence-state ceria nanoparticles could scavenge
these reactive oxygen intermediates and prevent degenerative reti-
nal disorders.198 Therefore, ceria nanoparticles may be used to treat
problems that cause blindness. Niu and co-workers showed that
ceria nanoparticles could protect cells in vivo in mice from lethal

Figure 6.22 Nanoceria particles provide pan-retinal protection against light dam-
age. The thickness of the outer nuclear layer (ONL) of the retina was measured
along the vertical meridian every 220 mm, starting at the optic nerve. Reprinted
with permission from Chen et al.198 Copyright 2006 Nature Publishing Group.

b1469_Ch-06.indd 337 4/8/2013 12:32:40 PM


b1469 Catalysis by Ceria and Related Materials

338 Z.-X. Li et al.

Figure 6.23 Ceria nanoparticles act as catalysts that mimic superoxide dismutase
(SOD) with a catalytic rate constant exceeding that determined for the enzyme
SOD. Reprinted with permission from Korsvik et al.200 Copyright 2007 the Royal
Chemical Society.

endoplasmic reticulum stress.199 The effect of ceria nanoparticles on


cardiac function was also assessed.
Self and co-workers used ceria nanoparticles as catalysts mimick-
ing superoxide dismutase (SOD).200 The polycrystalline 3–5 nm
ceria nanoparticles, with a Ce3+ concentration of 40%, had an excel-
lent catalytic rate constant even exceeding that of the enzyme SOD.
The mechanism should be further elucidated (Fig. 6.23).
Perez et al. prepared dextran-coated ceria nanoparticles by room
temperature hydrolysis in an ammonia aqueous solution in the pres-
ence of dextran T-10,201 which is bioactive monodispersed and water
soluble, and highly crystalline nanoparticles with pH-dependent anti-
oxidant activity. This product contained a ratio of Ce4+/Ce3+ of approx-
imately 3/2, which became 2/3 after adding some H2O2. After 10 days,
the nanoparticles were able to regenerate the original ratio when
under pH = 7.4, while there was no regeneration under pH = 4.
In order to improve selective antitumor therapy, Vincent et al. suc-
cessfully conjugated ceria nanoparticles to the tumor marker transfer-
rin.202 They also combined single molecule force spectroscopy (SMFS)

b1469_Ch-06.indd 338 4/8/2013 12:32:40 PM


b1469 Catalysis by Ceria and Related Materials

2D and 3D Ceria-Based Nanoarchitectures 339

and density functional theory (DFT) simulations to understand the


interaction between the transferrin in cancer cells and ceria nanopar-
ticles. SMFS studies illustrated that the positive potential of ceria nano-
particles increased with an increase of the transferrin adhesion to the
ceria nanoparticles. Binding energy values obtained from DFT calcu-
lations predict an increase in bond strength between the transferrin
and ceria nanoparticles on surface protonation and charge modifica-
tion (Fig. 6.24).
Several papers have placed much emphasis on the ability of
nanoceria to spontaneously switch from the oxidized (IV) to the
reduced (III) state, thus recycling its antioxidant ability. However,
the mechanism of the redox regeneration of ceria nanoparticles is
unclear. These works used the SOD-mimetic203,204 and the catalase-
mimetic205 activities together to provide a bio-related mechanism for
the regeneration of nanoceria. When nanoceria reduces superox-
ide, H2O2 is formed and Ce3+ is oxidized to Ce4+; then Ce4+ and H2O2
can react together to regenerate Ce3+ and oxidize H2O2 to O2. This
is a very elegant way of regenerating reduced nanoceria and elimi-
nating, in a sequential set of reactions, both superoxide and hydro-
gen peroxide. The stoichiometry requires two superoxides reduced
for each H2O2 oxidized (Fig. 6.25).193 Alternatively, a second mole-
cule of H2O2 may oxidize the reduced Ce3+,204 leading to the forma-
tion of Ce4+ and the reduction of H2O2 to H2O; in this case a true
catalase-like dismutation cycle has been created (Fig. 6.26).
Novel applications of nanoceria in the growing field of nano-
medicine appear every day, including tissue engineering and spe-
cific drug targeting. Recently, ceria nanoparticles were tested as
novel supports for stem cells culturing in vitro, and they could
improve the culturing of mesenchymal stem cells and cardiac pro-
genitors grown in a biodegradable polymer matrix, by promoting
their adhesion and proliferation, possibly via the antioxidant mech-
anism.206 The results are very promising for the pharmaceutical use
of nanoceria in clinical trials in the near future. For biomedical
applications, there needs to be a prompt safety evaluation. The tox-
icity of cerium in its Ce(III) form to Escherichia coli has been investi-
gated207 and a mechanism for the uptake of ceria by fibroblasts was

b1469_Ch-06.indd 339 4/8/2013 12:32:41 PM


b1469 Catalysis by Ceria and Related Materials

340 Z.-X. Li et al.

Figure 6.24 Atomic structural model and energy profile of a carboxyl ion of a
glycinate interacting with a ceria nanoparticle. (a) Side view of the relaxed configu-
ration of a negatively charged glycinate ion on a triply protonated facet of ceria
nanoparticle. (b), (c) Projection of the electron localization function (ELF) on
different planes passing through oxygen atoms of a protonated ceria nanoparticle
and a glycinate ion, revealing the weak and strong hydrogen bonds between them.
Color code blue (ELF = 0) and red (ELF = 1.0) are for the full absence and full
presence of an electron pair at an actual point of space, respectively. (d) Ground-
state energy profile. Blue curve: ground-state energy levels of a protonated ceria
nanoparticle (CNP) interacting with the carboxyl ion of a glycinate. Red curve:
sum of the ground-state energies of a protonated ceria nanoparticle and a glyci-
nate ion located at infinite distance. Cyan curve: binding energy between the
protonated ceria nanoparticle and a carboxyl ion of the glycinate evaluated as the
difference between the red and blue energy curves. ELF plots also reveal that all
the protons formed strong hydrogen bonds with oxygen atoms of cerium oxide.
Reprinted with permission from Vincent et al.202 Copyright 2009 the American
Chemical Society.

b1469_Ch-06.indd 340 4/8/2013 12:32:41 PM


b1469 Catalysis by Ceria and Related Materials

2D and 3D Ceria-Based Nanoarchitectures 341

Figure 6.25 Reaction mechanism for the oxidation of hydrogen peroxide by


nanoceria and the regeneration via reduction by superoxide. An oxygen vacancy
site on the nanoceria surface (1) presents a 2Ce4+ binding site for H2O2 (2), after
the release of protons and two-electron transfer to the two cerium ions (3) oxygen
is released from the now fully reduced oxygen vacancy site (4). Subsequently super-
oxide can bind to this site (5), and after the transfer of a single electron from one
Ce3+, and the uptake of two protons from the solution, H2O2 is formed (6) and
can be released. After repeating this reaction with a second superoxide molecule
(7) the oxygen vacancy site returns to the initial 2Ce4+ state (1). It is also possible
that the third Ce3+ indicated, which gives rise to the oxygen vacancy, could partici-
pate directly in the reaction mechanism. The square Ce–O matrix is shown here
only to illustrate the model and does not correspond to the actual spatial arrange-
ment of atoms in the crystal structure. Reprinted with permission from Celardo
et al.193 Copyright 2011 the Royal Chemical Society.

then proposed by Limbach et al.208 Recently, Thill et al. found that


direct spatial contact has to be made so that the cytotoxicity of CeO2
nanoparticles can be provoked on Escherichia coli.209 Not only “nega-
tive” but “positive” effects of ceria nanoparticles have been discov-
ered as well. Xia et al. carried out a mechanistic study to elucidate
the physicochemical characteristics of TiO2, ZnO, and CeO2, which

b1469_Ch-06.indd 341 4/8/2013 12:32:42 PM


b1469 Catalysis by Ceria and Related Materials

342 Z.-X. Li et al.

Figure 6.26 Reaction mechanism for the complete dismutation of hydrogen


peroxide. The oxidative half-reaction is identical to the sequence shown in Fig. 6.25
(1–4). The reductive half involves binding of H2O2 to the 2Ce3+ site (5), the uptake
of two protons and homolysis of the O–O bond with the transfer of electrons to
the two Ce3+ (6), and the release of water molecules to regenerate the initial Ce4+
site (1). This reaction sequence is analogous to the one found in catalases.
Reprinted with permission from Celardo et al.193 Copyright the Royal Chemical
Society 2011.

determine their toxic effects on biological processes, finding that


the antioxidant properties of CeO2 protect cells from oxidant injury
while the other two metal oxides induce cytotoxicity.210

6.3.4 Optical applications


Ceria also exhibits optical properties suitable for potential applica-
tions. Ceria has characteristics ideal for use as a broad-spectrum
inorganic sunscreen in personal care products: it is quite transparent
to visible light, but has excellent ultraviolet absorption ability. Thin
films of ceria have a high refractive index (n = 2.05), and dc

b1469_Ch-06.indd 342 4/8/2013 12:32:42 PM


b1469 Catalysis by Ceria and Related Materials

2D and 3D Ceria-Based Nanoarchitectures 343

dielectric constant, making them suitable for applications in optical,


electro-optical, microelectronic, and optoelectronic devices.
Patsalas et al. described the electron-beam evaporation (EBE)
growth of nanocrystalline (grains of 8–40 nm) ceria films. The
refractive index n and fundamental gap Eg of the nanocrystalline
ceria films were tailored independently by varying the substrate tem-
perature or using Ar1 ion beams during EBE growth.211 The film
composition strongly affected Eg, which varied from 2.8 to 2.0 eV.
The optical absorption below 3 eV and the Eg shift were attributed
to oxygen defect states and not to modifications in interband
transitions.
Photonic crystals, first introduced by Yablonovitch212 and John,213
have highly ordered structures with a periodically modulated refrac-
tive index (or dielectric constant), with periods typically of the
length scale of optical wavelengths. This periodicity may lead to the
formation of a photonic bandgap (PBG), a band of frequencies for
which light propagation in the photonic crystal is forbidden. The
position of the PBG can be adjusted by designing and modulating
the refractive index. Waterhouse et al. described the fabrication of
inverse opal CeO2 photonic crystal films and powders, through a
colloidal template sol-gel procedure with poly(methyl methacrylate)
(PMMA) spheres (Fig. 6.27).214,215 The films obtained had 3D
ordered macroporous structures and a photonic bandgap in the vis-
ible region. Filling the macropores of the CeO2 inverse opal with
solvents of different refractive index caused a shift in the position of
the photonic bandgap, with the magnitude of the shift being directly
proportional to the refractive index of the solvent. Refractive index
sensing with a sensitivity of n = 0.001 or better is achievable using
inverse opal CeO2 thin films.
Because of the high refractive index of ceria in the visible range
(2.05 eV), ceria appears natural on the skin without imparting an
excessively pale white look. Therefore, ceria is used as a UV-shielding
material for cosmetics. However, the photocatalytic activity of ceria,
which facilitates the generation of reactive oxygen species and has
raised safety concerns, must be suppressed for such applications.216–226
The activity can be reduced in various ways, such as doping ceria with

b1469_Ch-06.indd 343 4/8/2013 12:32:42 PM


b1469 Catalysis by Ceria and Related Materials

344 Z.-X. Li et al.

Figure 6.27 (a) SEM image of PMMA opal photonic crystal. (b) SEM image of ceria
inverse opal photonic crystal made from the PMMA template. Reprinted with per-
mission from Waterhouse et al.214 Copyright 2008 The American Chemical Society.

calcium, making the ceria into nanoparticles, as well as coating the


ceria nanoparticles with silica shells.227–229
Ceria only exhibits weak luminescence, therefore doping with
rare earths such as Eu3+ can enhance the visible emission required
for imaging. Leite and co-workers produced gadolinium-doped ceria
nanorods with a nominal composition of Gd0.2Ce0.8O2−x by hydro-
thermal treatment of a colloidal suspension of Gd-doped CeO2
nanocrystals in an aqueous solution (Fig. 6.28(a)).230 Wang et al.
reported that undoped ceria nanocrystals show a very weak emission
band peaking at 501 nm, which is remarkably enhanced by dop-
ing the ceria nanocrystals with lanthanide ions (Eu3+, Tb3+, Sm3+)
(Fig. 6.28(b), (c), and (d)).231 This large increase, located at 500 nm,
is not attributed to the characteristic emission of the lanthanide ions
(Eu3+, Tb3+, Sm3+) but to the presence of oxygen vacancies in the
CeO2 nanocrystal lattice.
Babu et al. synthesized Eu3+-doped ceria nanoparticles of 10 nm
by room temperature chemical precipitation and annealed at differ-
ent temperatures to study the effects on luminescence.232 The num-
ber of Ce3+ ions in the ceria nanoparticles decreased after annealing.
The emission intensity varied with the wavelength of excitation and
observed transitions indicated the presence of Eu3+ in different sym-
metry environments.

b1469_Ch-06.indd 344 4/8/2013 12:32:42 PM


b1469 Catalysis by Ceria and Related Materials

2D and 3D Ceria-Based Nanoarchitectures 345

Figure 6.28 HRTEM image of (a) Gd-doped CeO2 nanorod (reprinted with per-
mission from Godinho et al.230 Copyright 2008 the American Chemical Society) and
(b) CeO2/Eu3+ nanocrystals. Photographs of (c) CeO2 nanocrystals and (d) CeO2/
Eu3+ nanocrystals (in a hexane solution). (e) Excitation and emission spectra of
CeO2 (red dashed line) and CeO2/Eu3+ (black solid line) nanocrystals in a hexane
solution. Reprinted with permission from Wang et al.231 Copyright 2007 the
American Chemical Society.

6.4 Conclusions and Outlook


There has been significant progress in the nanoscale controlled
synthesis of ceria-based materials in recent decades. Different shapes
and sizes of ceria have been produced with various synthesis
methods developed by many research groups.233 However, it is not
an exaggeration to say that the chemical synthesis of ceria-based
nanomaterials remains a technique rather than a science because
the atomistic details for the evolution from a precursor to the final
well-defined nanostructures are far from being understood. The

b1469_Ch-06.indd 345 4/8/2013 12:32:42 PM


b1469 Catalysis by Ceria and Related Materials

346 Z.-X. Li et al.

production of high-quality ceria-based nanomaterials in a controlled


manner and a detailed understanding of the synthetic mechanisms
are still challenges to be faced in the coming years. The develop-
ment of in situ monitoring techniques opens up opportunities for
revealing the mechanisms of the crystal nucleation and growth pro-
cesses. Electrical measurements developed by our group have shed
light on these processes and other measurement techniques such as
optical, spectral, and so on are required urgently.
Meanwhile, simulation has become an efficient way for us to
understand the formation of nanocrystals at an atomic level. The
MD simulation of crystal growth, the surface effect induced phase,
and the control of morphology are essential for revealing structural
characteristics, formation mechanisms, and shape control factors.
Our group has started to simulate the surfactant-modified shape
evolution of ceria nanomaterials. Only an integration of experimen-
tal evidence and theoretical modeling can give us a scientific under-
standing of material synthesis and guide synthesis for applications.
A pre-designed synthesis with a precisely adjusted Ce4+/Ce3+
ratio, the defect content, is required for special applications in
catalysis, such as sensors and so on. Catalysis is now the most impor-
tant use of ceria-based nanomaterials. However, the intrinsic Ce4+/
Ce3+ change properties of ceria suggest that it can also fully use its
capabilities in biomedicine. Pioneering research has revealed that
ceria nanoparticles could be effective in inhibiting the progression
of reactive oxygen intermediates, which induce cell death, and so
could be useful in therapy for diseases that cause blindness.
Nanoceria was also found to be effective for spinal-cord repair, treat-
ing other diseases of the central nervous system, and for protecting
normal human cells from irradiation. The potential of vacancy-
engineered ceria in disease therapy and anti-aging are sufficiently
fascinating to push us towards further research in this field. Novel
applications of nanoceria in the growing field of nanomedicine
appear every day, including tissue engineering and specific drug
targeting. It was recently shown that nanoceria improves the cultur-
ing of mesenchymal stem cells and cardiac progenitors grown in a
biodegradable polymer matrix, by promoting their adhesion and

b1469_Ch-06.indd 346 4/8/2013 12:32:43 PM


b1469 Catalysis by Ceria and Related Materials

2D and 3D Ceria-Based Nanoarchitectures 347

proliferation, possibly via the antioxidant mechanism. Moreover,


nanoceria was successfully conjugated to the tumor marker transfer-
rin to improve selective antitumor therapy.
Further demands include environmentally friendly, cost-effec-
tive, high-throughput, and reliable nanocrystals with a precisely
controlled shape and size and desired surface states. In many cases,
especially when a surfactant or template is used in synthesis, the
surface of a ceria-based nanoparticle is covered with organic species,
which is a disadvantage for further application in catalysis. Another
problem in synthesis is that due to the different nucleation growth
and hydrolysis behavior of cerium and zirconium species, a well-
controlled synthesis of ceria-zirconia solid-solution nanocrystals with
homogeneous compositions, tunable shapes, and uniform sizes is
rarely achieved. There is also the potential for further improve-
ments in the shape-controlled synthesis of other ceria-based com-
posites. New strategies need to be exploited to produce new types of
assembly structure. Understanding the collective properties of an
assembly structure is critical for controlling assembly and for devel-
oping new materials and devices from the nanoscale building
blocks. For example, for understanding the relationship between
the facet and catalytic activity, uniform ceria nanoparticles are
expected to arrange to expose the specific facet by using the assem-
bly technique. Multicomponent assembly is an efficient way to
realize multifunction properties; ceria nanocrystals and other
nanocrystals can be designed in various ways, for example, different
components can be linked together in a binary manner or with
more building blocks, for applications such as the magnetic separa-
tion of a catalyst or other new nanodevices. The dimensions and
control of the morphology of the assembled ceria-based materials
are also important for determining the collective properties.

References
1. Iijima, S., Nature, 354 (1991) 56–58.
2. Delgado, J.L., Herranz, M., Martin, N., J. Mater. Chem., 18 (2008)
1417–1426.

b1469_Ch-06.indd 347 4/8/2013 12:32:43 PM


b1469 Catalysis by Ceria and Related Materials

348 Z.-X. Li et al.

3. Thess, A., Lee, R., Nikolaev, P., Dai, H., Petit, P., Robert, J., Xu, C., Lee,
Y.H., Kim, S.G., Rinzler, S.G., Colbert, D.T., Scuseria, G.E., Tomanek, D.,
Fischer, J.E., Smalley, R.E., Science, 273 (1996) 483–487.
4. Tans, S.J., Verschueren, A.R.M., Dekker, C., Nature, 393 (1998) 49–52.
5. Deheer, W.A., Chatelain, A., Ugarte, D., Science, 270 (1995) 1179–1180.
6. Treacy, M.M.J., Ebbesen, T.W., Gibson, J.M., Nature, 381 (1996) 678–680.
7. Fan, S.S., Chapline, M.G., Franklin, N.R., Tombler, T.W., Cassell, A.M.,
Dai, H., Science, 283 (1999) 512–514.
8. Novoselov, K.S., Geim, A.K., Morozov, S.V., Jiang, D., Katsnelson, M.I.,
Grigorieva, I.V., Dubonos, S.V., Firsov, A.A., Nature, 438 (2005)
197–200.
9. Geim, A.K., Novoselov, K.S., Nat. Mater., 3 (2007) 183–191.
10. Alivisatos, A.P., Science, 271 (1996) 933–937.
11. Murray, C.B., Norris, D.J., Bawendi, M.G., J. Am. Chem. Soc., 115
(1993) 8706–8715.
12. Klimov, V.I., Mikhailovsky A.A., Xu, S., Malko, A., Hollingsworth, J.A.,
Leatherdale, C.A., Eisler, H.-J., Bawendi, M.G., Science, 290 (2000)
314–317.
13. Bruchez Jr., M., Moronne, M., Gin, P., Weiss, S., Alivisatos, A.P., Science,
281 (1998) 2013–2016.
14. Ozgur, U., Alivov, Y.I., Liu, C., Teke, A., Reshchikov, M.A., Dogan, S.,
Avrutin, V., Cho, S.-J., Morkoc, H., J. Appl. Phys., 98 (2005) 041301–
041301-103.
15. Wang, Z.L., Song, J.H., Science, 312 (2006) 242–246.
16. Kelly, K.L., Coronado, E., Zhao, L.L., Schatz, G.C., J. Phys. Chem. B,
107 (2003) 668–677.
17. Homola, J., Yee, S.S., Gauglitz, G., Sens. Act. B, 54 (1999) 3–15.
18. Nie, S.M., Emery, S.R., Science, 275 (1997) 1102–1106.
19. Kneipp, K., Wang, Y., Kneipp, H., Perelman, L.T., Itzkan, I.,
Ramachandra, R.D., Feld, M.S., Phys. Rev. Lett., 78 (1997) 1667–1670.
20. Xia, Y.N., Yang, P.D., Sun, Y.G., Wu, Y., Mayers, B., Gates, B., Yin, Y.,
Kim, F., Yan, H., Adv. Mater., 15 (2003) 353–389.
21. Trovarelli, A., ed. in Catalysis by Ceria and Related Materials, Imperial
College Press, London, (2002).
22. Kašpar, J., Fornasiero, P., Graziani, M., Catal. Today, 50 (1999)
285–298.

b1469_Ch-06.indd 348 4/8/2013 12:32:43 PM


b1469 Catalysis by Ceria and Related Materials

2D and 3D Ceria-Based Nanoarchitectures 349

23. Kašpar, J., Graziani, M., Fornasiero, P., in Handbook on the Physics and
Chemistry of Rare Earths: The Role of Rare Earths in Catalysis, eds. K.A.
Gschneidner Jr., L. Eyring, Elsevier Science B.V., Amsterdam, (2000),
pp. 159–267.
24. Steele, B.C.H., Nature, 400 (1999) 619–621.
25. Steele, B.C.H., J. Mater. Sci., 36 (2001) 1053–1068.
26. Steele, B.C.H., Heinzel, A., Nature, 414 (2001) 345–352.
27. Li, R.X., Yabe, S., Yamashita, M., Momose, S., Yoshida, S., Yin, S.,
Sato, T., Solid State Ionics, 151 (2002) 235–241.
28. Tsunekawa, S., Sivamohan, R., Ohsuna, T., Mater. Sci. Forum, 315–317
(1999) 439–445.
29. Wu, L.J., Wiesmann, H.J., Moodenbaugh, A.R., Kile, R.F., Zhu, Y.,
Welch, D.O., Suenaga, M., Phys. Rev. B, 69 (2004) 125415.
30. Tsunekawa, S., Fukuda, T., Kasuya, A., Surf. Sci., 457 (2000) L437–440.
31. Carrettin, S., Concepcion, P., Corma, A., Lopez Nieto, J.M., Puntes, V.F.,
Angew. Chem., Int. Ed., 43 (2004) 2538–2540.
32. Tschope, A., Sommer, E., Birringer, R., Solid State Ionics, 139 (2001)
255–265.
33. Liao, L., Mai, H.X., Yuan, Q., Lu, H.B., Li, J.C., Liu, C., Yan, C.H.,
Shen, Z.X., Yu, T., J. Phys. Chem. C, 112 (2008) 9061–9065.
34. Chen, L., Fleming, P., Morris, V., Holmes, J.D., Morris, M.A., J. Phys.
Chem. C, 114 (2010) 12909–12919.
35. Yamashita, M., Kameyama, K., Yabe, S., Yoshhida, S., Fujishiro, Y.,
Kawai, T., Sato, T., J. Mater. Sci., 37 (2002) 683–687.
36. Jin, H.Y., Wang, N., Xu, L., Hou, S., Mater. Lett., 64 (2010) 1254–1256.
37. Zhou, X.D., Huebner, W., Anderson, H.U., Appl. Phys. Lett., 80 (2002)
3814–3816.
38. Chen, J.C., Chen, W.C., Tien, Y.C., Shih, C.J., J. Alloys Compd., 496
(2010) 364–369.
39. Matijevic, E., Hsu, W.P., J. Colloid Interface Sci., 118 (1987) 506–523.
40. Tsai, M.S., Mat. Sci. Eng. B-Solid, 110 (2004) 132–134.
41. Chen, F.J., Cao, Y.L., Jia, D.Z., Appl. Surf. Sci., 257 (2011) 9226–9231.
42. Wang, G.F., Mu, Q.Y., Chen, T., Wang, Y., J. Alloys Compd., 493 (2010)
202–207.
43. Liu, J., Wang, L.N., Wu, H.M., Mo, L.Y., Lou, H., Zheng, X.M., React.
Kinet. Catal. Lett., 98 (2009) 311–318.

b1469_Ch-06.indd 349 4/8/2013 12:32:43 PM


b1469 Catalysis by Ceria and Related Materials

350 Z.-X. Li et al.

44. Chaudhary, Y.S., Panigrahi, S., Nayak, S., Satpati, B., Bhattacharjee, S.,
Kulkarni, N., J. Mater. Chem., 20 (2010) 2381–2385.
45. Zhang, D.S., Niu, F.H., Li, H.R., Shi, L.Y., Fang, J.H., Powder Technol.,
207 (2011) 35–41.
46. Hakuta, Y., Onai, S., Terayama, H., Adschiri, T., Arai, K., J. Mater. Sci.
Lett., 17 (1998) 1211–1213.
47. Hirano, M., Kato, E., J. Am. Ceram. Soc., 82 (1999) 786–788.
48. Wu, N.C., Shi, E.W., Zheng, Y.Q., Li, W.J., J. Am. Ceram. Soc., 85 (2002)
2462–2468.
49. Hirano, M., Kato, E., J. Mater. Sci. Lett., 15 (1996) 1249–1250.
50. Zhou, Y.C., Rahaman, M.N., J. Mater. Res., 8 (1993) 1680–1686.
51. Lakhwani, S., Rahaman, M.N., J. Mater. Res., 14 (1999) 1455–1461.
52. Hirano, M., Fukuda, Y., Iwata, H., Hotta, Y., Inagaki, M., J. Am. Ceram.
Soc., 83 (2000) 1287–1289.
53. Wang, Z.L., Feng, X.D., J. Phys. Chem. B, 107 (2003) 13563–13566.
54. Fu, C., Li, R.X., Zhang, Y., Int. J. Mod. Phys. B, 24 (2010) 3230–3235.
55. Yin, L.X., Wang, Y.Q., Pang, G.S., Koltypin, Y., Gedankan, A., J. Colloid
Interface Sci., 246 (2002) 78–84.
56. Xu, H.R., Gao, L., Gu, H.C., Guo, J.K., Yan, D.S., J. Am. Ceram. Soc., 85
(2002) 139–144.
57. Verma, A., Karar, N., Bakhshi, A.K., Chander, H., Shivaprasad, S.M.,
Agnihotri, S.A., J. Nanopart. Res., 9 (2007) 317–322.
58. Chu, X., Chung, W.I., Schmidt, L.D., J. Am. Ceram. Soc., 76 (1993)
2115–2118.
59. Oh, M.H., Nho, J.S., Cho, S.B., Lee, J.S., Singh, R.K., Mater. Chem.
Phys., 124 (2010) 134–139.
60. Fu, C., Li, R.X., Tang, Q., Li, C.Q., Yin, S., Sato, T., Res. Chem.
Intermediat., 37 (2011) 319–327.
61. Masui, T., Fujiwara, K., Machida, K., Adachi, G.Y., Sakata, T., Mori, H.,
Chem. Mater., 9 (1997) 2197–2204.
62. Zhou, H.P., Zhang, Y.W., Si, R., Zhang, L.S., Song, W.G., Yan, C.H.,
J. Phys. Chem. C, 112 (2008) 20366–20374.
63. Goharshadi, E.K., Samiee, S., Nancarrow, P., J. Colloid Interface Sci., 356
(2011) 473–480.
64. Wilson, M.S., Delariva, A., Garzon, F.H., J. Mater. Chem., 21 (2011)
7418–7424.

b1469_Ch-06.indd 350 4/8/2013 12:32:43 PM


b1469 Catalysis by Ceria and Related Materials

2D and 3D Ceria-Based Nanoarchitectures 351

65. Wu, Z.L., Li, M.J., Howe, J., Meyer Ill, H.M., Overbury, S.H., Langmuir,
26 (2010) 16595–16606.
66. Lv, J.G., Shen, Y., Peng, L.M., Guo, X.F., Ding, W.P., Chem. Commun.,
46 (2010) 5909–5911.
67. Yang, Z.X., Woo, T.K., Baudin, M., Hermansson, K., J. Chem. Phys., 120
(2004) 7741–7749.
68. Skorodumova, N.V., Baudin, M., Hermansson, K., Phys. Rev. B, 69
(2004) 075401.
69. Conesa, J.C., Surf. Sci., 339 (1995) 337–352.
70. Sayle, T.X.T., Parker, S.C., Sayle, D.C., Chem. Commun., (2004)
2438–2439.
71. Yang, S.W., Gao, L., J. Am. Chem. Soc., 128 (2006) 9330–9331.
72. Taniguchi, T., Katsumata, K., Omata, S., Okada, K., Matsuhita, N.,
Cryst. Growth & Des., 11 (2011) 3754–3760.
73. Wu, Q., Zhang, F., Xiao, P., Tao, H.S., Wang, X.Z., Hu, Z., Lu, Y.N.,
J. Phys. Chem. C, 112 (2008) 17076–17080.
74. Qiu, H.L., Chen, G.Y., Fan, R.W., Cheng, C., Hao, S.W., Chen, D.Y.,
Yang, C.H., Chem. Commun., 47 (2011) 9648–9650.
75. Lin, K.S., Chowdhury, S., Int. J. Mol. Sci., 11 (2010) 3226–3251.
76. Yu, K.L., Ruan, G.L., Ben, Y.H., Zou, J.J., Mat. Sci. Eng. B-Solid, 139
(2007) 197–200.
77. Yang, T., Xia, D.G., Mater. Chem. Phys., 123 (2010) 816–820.
78. Li, Y.H., Ding, J., Chen, J.F., Xu, C.L., Wei, B.Q., Liang, J., Wu, D.H.,
Mater. Res. Bull., 37 (2002) 313–318.
79. Zhang, D.S., Fu, H.X., Shi, L.Y., Fang, J.H., Li, Q., J. Solid State Chem.,
180 (2007) 654–660.
80. Wei, J.Q., Ding, J., Zhang, X.F., Wu, D.H., Wang, Z.C., Luo, J.B.,
Wan, K.L., Mater. Lett., 59 (2005) 322–325.
81. Fang, J.H., Cao, Z.Y., Zhang, D.S., Shen, X., Ding, W.Z., Shi, L.Y.,
J. Rare Earth., 26 (2008) 153–157.
82. Zhang, D.S., Shi, L.Y., Fu, H.X., Fang, J.H., Carbon, 44 (2006)
2853–2855.
83. Fu, H.X., Zhang, D.S., Shi, L.Y., Fang, J.H., Chem. J. Chinese U., 28
(2007) 617–620.
84. Zhang, D.S., Pan, C.S., Shi, L.Y., Huang, L., Fang, J.H., Fu, H.X.,
Micropor. Mesopor. Mat., 117 (2009) 193–200.

b1469_Ch-06.indd 351 4/8/2013 12:32:43 PM


b1469 Catalysis by Ceria and Related Materials

352 Z.-X. Li et al.

85. Zhang, L.S., Jiang, L.Y., Chen, C.Q., Li, W., Song, W.G., Guo, Y.G.,
Chem. Mater., 22 (2010) 414–419.
86. Chen, Z.G., Chen, F., Li, X.Z., Lu, X.W., Ni, C.Y., Zhao, X.B., J. Rare
Earth, 28 (2010) 566–570.
87. Boehme, M., Fu, G., Ionescu, E., Ensinger, W., Nanotechnology, 22
(2011) 065602.
88. Vantomme, A., Yuan, Z.Y., Du, G.H., Su, B.L., Langmuir, 21 (2005)
1132–1135.
89. Yang, Y.F., Jin, Y.Z., He, H.P., Ye, Z.Z., CrystEngComm, 12 (2010)
2663–2665.
90. Yu, T.Y., Joo, J., Park, Y.I., Hyeon, T., Angew. Chem. Int. Ed., 44 (2005)
7411–7414.
91. Huo, Z.Y., Tsung, C.K., Huang, W.Y., Zhang, X.F., Yang, P.D., Nano
Lett., 8 (2008) 2041–2044.
92. Lu, X.M., Yavuz, M.S., Tuan, H.Y., Korgel, B.A., Xia, Y., J. Am. Chem.
Soc., 130 (2008) 8900–8901.
93. Han, W.Q., Wu L.J., Zhu, Y.M., J. Am. Chem. Soc., 127 (2005)
12814–12815.
94. Chen, H.I., Chang, H.Y., Solid State Commun., 133 (2005) 593–598.
95. Mai, H.X., Sun, L.D., Zhang, Y.W., Si, R., Feng, W., Zhang, H.P., Liu,
H.C., Yan, C.H., J. Phys. Chem. B, 109 (2005) 24380–24385.
96. Pan, C.S., Zhang, D.S., Shi, L.Y., Fang, J.H., Eur. J. Inorg. Chem., (2008)
2429–2436.
97. Wang, W., Howe, J.Y., Li, Y.A., Qiu, X.F., Joy, D.C., Paranthaman, M.P.,
Doktycz, M.J., Gu, B.H., J. Mater. Chem., 20 (2010) 7776–7781.
98. Chen, G.Z., Xu, C.X., Song, X.Y., Zhao, W., Ding, Y., Sun, S.X., Inorg.
Chem., 47 (2008) 723–728.
99. Zhou, K.B., Yang, Z.Q., Yang, S., Chem. Mater., 19 (2007) 1215–1217.
100. Miao, J.J., Wang, H., Li, Y.R., Zhu, J.M., Zhu, J.J., J. Cryst. Growth, 281
(2005) 525–529.
101. Kuiry, S.C., Patil, S.D., Deshpande, S., Seal, S., J. Phys. Chem. B, 109
(2005) 6936–6939.
102. Ge, M., Guo, C.S., Li, L., Zhang, B.Q., Feng, Y.C., Wang, Y.Q., Mater.
Lett., 63 (2009) 1269–1271.
103. Sun, Z.Y., Zhang, H.Y., An, G.M., Yang, G.Y., Liu, Z.M., J. Mater. Chem.,
20 (2010) 1947–1952.

b1469_Ch-06.indd 352 4/8/2013 12:32:43 PM


b1469 Catalysis by Ceria and Related Materials

2D and 3D Ceria-Based Nanoarchitectures 353

104. Martin, P., Parker, S.C., Sayle, D.C., Watson, G.W., Nano Lett., 7 (2007)
543–546.
105. Zhang, D.E., Zhang, X.J., Ni, X.M., Song, J.M., Zheng, H.G.,
Chemphyschem, 7 (2006) 2468–2470.
106. Zhang, D.S., Niu, F.H., Yan, T.T., Shi, L.Y., Du, X.J., Fang, J.H., Appl.
Surf. Sci., 257 (2011) 10161–10167.
107. Sun, C.W., Chen, L.Q., Eur. J. Inorg. Chem., (2009) 3883–3887.
108. Zhang, D.E., Wu, W., Ni, X.J., Cao, X.Y., Zhang, X.B., Xu, X.Y., Li, S.Z.,
Han, G.Q., Ying, A.L., Tong, Z.W., J. Mater. Sci., 44 (2009) 3344–3348.
109. Wang, D.Y., Kang, Y.J., Doan-Nguyen, V., Chen, J., Kungas, R.,
Wider, N.L., Bakhmutsky, K., Gorte, R.J., Murray, C.B., Angew. Chem.,
Int. Ed., 50 (2011) 4378–4381.
110. Minamidate, Y., Yin, S., Sato, T., Mater. Chem. Phys., 123 (2010)
516–520.
111. Chen, J.J., Zhong, S.L., Liu, Q.Y., Wang, Y.L., Wang, S.P., Xu, R.,
Luo, L.F., Wang, S.J., Powder Technol., 197 (2010) 136–139.
112. Zhang, D.S., Huang, L., Zhang, J.P., Shi, L.Y., J. Mater. Sci., 43 (2008)
5647–5650.
113. Wang, Z.L., Li, G.R., Ou, Y.N., Feng, Z.P., Qu, D.L., Tong, Y.X., J. Phys.
Chem. C, 115 (2011) 351–356.
114. Zhou, H.P., Zhang, Y.W., Mai, H.X., Sun, X., Liu, Q., Song, W.G.,
Yan, C.H., Chem. Eur. J., 14 (2008) 3380–3390.
115. Li, J.F., Lu, G.Z., Li, H.F., Wang, Y.Q., Guo, Y., Guo, Y.L., J. Colloid
Interface Sci., 360 (2011) 93–99.
116. Kempaiah, D.M., Yin, S., Sato, T., CrystEngComm, 13 (2011) 741–746.
117. Zhang, D.S., Yan, T.T., Pan, C.S., Shi, L.Y., Zhang, J.P., Mater. Chem.
Phys., 113 (2009) 527–530.
118. Guo, Z.Y., Jian, F.F., Du, F.L., Scripta Mater., 61 (2009) 48–51.
119. Strandwitz, N.C., Stucky, G.D., Chem. Mater., 21 (2009) 4577–4582.
120. Sun, X.M., Liu, J.F., Li, Y.D., Chem. Eur. J., 12 (2006) 2039–2047.
121. Cao, C.Y., Cui, Z.M., Chen, C.Q., Song, W.G., Cai, W., J. Phys. Chem. C,
114 (2010) 9865–9870.
122. Yang, Z.J., Han, D.Q., Ma, D.L., Liang, H., Liu, L., Yang, Y.Z., Cryst.
Growth & Des., 10 (2010) 291–295.
123. Chen, G.Z., Zhu, F.F., Sun, X.A., Sun, S.X., Chen, R.P., CrystEngComm,
13 (2011) 2904–2908.

b1469_Ch-06.indd 353 4/8/2013 12:32:44 PM


b1469 Catalysis by Ceria and Related Materials

354 Z.-X. Li et al.

124. Zhang, Y.J., Cheng, T., Hu, Q.X., Fang, Z., Han, K., J. Mater. Res., 22
(2007) 1472–1478.
125. Chen, G.Z., Xu, C.X., Song, X.Y., Xu, S.L., Ding, Y., Sun, S.X., Cryst.
Growth & Des., 8 (2008) 4449–4453.
126. Lu, X.W., Li, X.Z., Chen, F., Ni, C.Y., Chen, Z.G., J. Alloys Compd., 476
(2009) 958–962.
127. Motte, L., Billoudet, F., Pileni, M.P., J. Phys. Chem., 99 (1995)
16425–16429.
128. Murray, C.B., Kagan, C.R., Bawendi, M.G., Science, 270 (1995)
1335–1338.
129. Corma, A., Atienzar, P., Garcia, H., Chane-Ching, J.Y., Nat. Mater., 3
(2004) 394–397.
130. Lyons, D.l M., Ryan, K.M., Morris, M.A., J. Mater. Chem., 12 (2004)
1207–12012.
131. Yu, J.C., Zhang, L., Lin, J., J. Colloid Interf. Sci., 260 (2003) 240–243.
132. Si, R., Zhang, Y.W., Xiao, C.X., Li, S.J., Lin, B.X., Kou, Y., Yan, C.H.,
Phys. Chem. Chem. Phys., 6 (2004) 1056–1063.
133. Yuan, Q., Liu, Q., Song, W.G., Feng, W., Pu, W.L., Sun, L.D., Zhang, Y.W.,
Yan, C.H., J. Am. Chem. Soc., 129 (2007) 6698–6699.
134. Li, Z.X., Li, L.L., Yuan, Q., Feng, W., Xu, J., Sun, L.D., Song, W.G.,
Yan, C.H., J. Phys. Chem. C, 112 (2008) 18405–18411.
135. Li, L.L., Xu, J., Yuan, Q., Li, Z.X., Song, W.G., Yan, C.H., Small, 5
(2009) 2730–2737.
136. Chane-Ching, J.Y., Cobo, F., Aubert, D., Harvey, H.G., Airiau, M.,
Corma, A., Chem. Eur. J., 11 (2005) 979–987.
137. Deshpande, A.S., Pinna, N., Smarsly, B., Antonietti, M., Niederberger, M.,
Small, 1 (2005) 313–316.
138. Si, R., Zhang, Y.W., You, L.P., Yan, C.-H., J. Phys. Chem. B, 110 (2006)
5994–6000.
139. Huo, Z.Y., Chen, C., Liu, X.W., Chu, D.R., Li, H.H., Peng, Q., Li, Y.D.,
Chem. Commun., 32 (2008) 3741–3743.
140. Sayle, D.C., Seal, S., Wang, Z.W., Mangili, B.C., Price, D.W., Karakoti,
A.S., Kuchibhatla, S.V.T.N., Hao, Q., Mobus, G., Xu, X.J., Sayle, T.X.T.,
ACS Nano, 2 (2008) 1237–1251.
141. Karakoti, A.S., Kuchibhatla, S.V.N.T., Baer, D.R., Thevuthasan, S.,
Sayle, D.C., Seal, S., Small, 4 (2008) 1210–1216.

b1469_Ch-06.indd 354 4/8/2013 12:32:44 PM


b1469 Catalysis by Ceria and Related Materials

2D and 3D Ceria-Based Nanoarchitectures 355

142. Kuchibhatla, S.V.N.T., Karakoti, A.S., Sayle, D.C., Heinrich, H., Seal, S.,
Cryst. Growth Des., 9 (2009) 1614–1620.
143. Kuchibhatla, S.V.N.T., Karakoti, A.S., Seal, S., Nanotechnology, 18
(2007) 075303.
144. Zhou, X.D., Huebner, W., Appl. Phys. Lett., 79 (2001) 3512.
145. Wang, Z.W., Saxena, S.K., Pischedda, V., Liermann, H.P., Zha, C.S.,
Phys. Rev. B, 64 (2001) 012102.
146. Wang, Z.W., Zhao, Y.S., Schiferl, D., Zha, C.S., Downs, R.T., Appl. Phys.
Lett., 85 (2004) 124–126.
147. Tsunekawa, S., Ishikawa, K., Li, Z.Q., Kawazoe, Y., Kasuya, A., Phys. Rev.
Lett., 85 (2000) 3440–3443.
148. Zhang, Y.W., Si, R., Liao, C.S., J. Phys. Chem. B, 107 (2003)
10159–10167.
149. Tsunekawa, S., Fukuda, T., Kasuya, A., J. Appl. Phys., 87 (2000)
1318–1321.
150. Spanier, J.E., Robinson, R.D., Zhang, F., Chan, S.W., Herman, I.P.,
Phys. Rev. B, 64 (2001) 245407.
151. Sayle, T.X.T., Parker, S.C., Catlow, C.R.A., Surf. Sci., 316 (1994)
329–336.
152. Sayle, T.X.T., Parker, S.C., Catlow, C.R.A., J. Chem. Soc., Chem. Commun.,
14 (1992) 977–978.
153. Wu, L.J., Wiesmann, H.J., Moodenbaugh, A.R., Klie, R.F., Zhu, Y.M.,
Welch, D.O., Suenaga, M., Phys. Rev. B, 69 (2004) 125415.
154. Murray, E.P., Tsai, T., Barnett, S.A., Nature, 400 (1999) 649–651.
155. Park, S.D., Vohs, J.M., Gorte, R.J., Nature, 404 (2000) 265–267.
156. Flytzani-Stephanopoulos, M., Sakbodin, M., Wang, Z., Science, 312
(2006) 1508–1510.
157. Azad, S., Marina, O.A., Wang, C.M., Appl. Phys. Lett., 86 (2005) 131906.
158. Sholklapper, T.Z., Kurokawa, H., Jacobson, C.P., Visco, S.J., De Jonghe,
L.C., Nano Lett., 7 (2007) 2136–2141.
159. Bernal, S., Calvino, J.J., Cauqui, M.A., Catal. Today, 50 (1999)
175–206.
160. Fu, Q., Weber, A., Flytzani-Stephanopoulos, M., Catal. Lett., 77 (2001)
87–95.
161. Fu, Q., Kudriavtseva, S., Saltsburg, H., Flytzani-Stephanopoulos, M.,
Chem. Eng. J., 93 (2003) 41–53.

b1469_Ch-06.indd 355 4/8/2013 12:32:44 PM


b1469 Catalysis by Ceria and Related Materials

356 Z.-X. Li et al.

162. Fu, Q., Saltsburg, H., Flytzani-Stephanopoulos, M., Science, 301 (2003)
935–938.
163. Liu, W., Flytzani-Stephanopoulos, M., J. Catal., 153 (1995) 304–316.
164. Liu, W., Flytzani-Stephanopoulos, M., J. Catal., 153 (1995) 317–332.
165. Fu, Q., Deng, W.L., Saltsburg, H., Flytzani-Stephanopoulos, M., Appl.
Catal. B, 56 (2005) 57–68.
166. Deng, W.L., Flytzani-Stephanopoulos, M., Angew. Chem., Int. Ed., 45
(2006) 2285–2289.
167. Corma, A., Domine, M.E., Chem. Commun., (2005) 4042–4044.
168. Guzman, J., Carrettin, S., Fierro-Gonzalez, J.C., Hao, Y.L., Gates, B.C.,
Corma, A., Angew. Chem., Int. Ed., 44 (2005) 4778–4781.
169. Carrettin, S., Corma, A., Iglesias, M., Sanchez, A., Appl. Catal. A, 291
(2005) 247–252.
170. Sedmak, G., Hočevar, S., Levec, J., J. Catal., 213 (2003) 135–150.
171. Avgouropoulos, G., Ioannides, T., Appl. Catal. A, 244 (2003) 155–167.
172. Ratnasamy, P., Srinivas, D., Satyanarayana, C.V.V., Manikandan, P.,
Senthil Kumaran, R.S., Sachin, M., Shetti, V.N., System J. Catal., 221
(2005) 455–465.
173. Murugan, B., Ramaswamy, A.V., J. Am. Chem. Soc., 129 (2007)
3062–3063.
174. Concepcion, P., Corma, A., Silvestre-Albero, J., Franco, V., Chane-
Ching, J.Y., J. Am. Chem. Soc., 126 (2005) 5523–5532.
175. Guzman, J., Carrettin, S., Corma, A., J. Am. Chem. Soc., 127 (2004)
3286–3287.
176. Mamontov, E., Egami, T., Brezny, R., Koranne, M., Tyagi, S., J. Phys.
Chem. B, 104 (2000) 11110.
177. Si, R., Zhang, Y.W., Li, S.J., Lin, B.X., Yan, C.H., J. Phys. Chem. B, 108
(2004) 12481–12488.
178. Zhou, K.B., Wang, X., Sun, X.M., Peng, Q., Li, Y.D., J. Catal., 229
(2005) 206–212.
179. Si, R., Flytzani-Stephanopoulos, M., Angew. Chem., Int. Ed., 47 (2008)
2884–2887.
180. Liu, P., Rodriguez, J.A., J. Chem. Phys., 126 (2007) 164705.
181. Abad, A., Almela, C., Corma, A., Garcia, H., Chem. Commun., 30
(2006) 3178–3180.
182. Abad, A., Corma, A., García, H., Chem. Eur. J., 14 (2008) 212–222.

b1469_Ch-06.indd 356 4/8/2013 12:32:44 PM


b1469 Catalysis by Ceria and Related Materials

2D and 3D Ceria-Based Nanoarchitectures 357

183. Zhou, H.P., Wu, H.S., Shen, J., J. Am. Chem. Soc., 132 (2010)
4998–4999.
184. Fu, X.Q., Wang, C., Yu, H.C., Wang, Y.G., Wang, T.H., Nanotechnology,
18 (2007) 145503.
185. Barreca, D., Gasparotto, A., Maccato, C., Maragno, C., Tondello, E.,
Langmuir, 22 (2006) 8639–8641.
186. Barreca, D., Gasparotto, A., Maccato, C., Maragno, C., Tondello, E.,
Comini, E., Sberveglieri, G., Nanotechnology, 18 (2007) 125502.
187. Ansari, A.A., Solanki, P.R., Malhotra, B.D., Appl. Phys. Lett., 92 (2008)
263901.
188. Saha, S., Arya, S.K., Singh, S.P., Sreenivas, K., Malhotra, B.D., Gupta, V.,
Biosens. Bioelect., 24 (2009) 2040–2045.
189. Guo, Z.Y., Du, F.L., Li, G.C., Cui, Z.L., Cryst. Growth & Des., 8 (2008)
2674–2677.
190. Zhang, W., Yang, T., Zhuang, X.M., Guo, Z.Y., Jiao, K., Biosens.
Bioelectron., 24 (2009) 2417–2422.
191. Ye, Q., Gao, Q., Zhang, X.R., Xu, B.Q., Catal. Comm., 7 (2006)
589–592.
192. Xuan, Y.L., Hu, J., Xu, K.L., Hou, X.D., Lu, Y., Sens. Act. B, 136 (2009)
218–223.
193. Celardo, I., Pedersen, J.Z. Traversab, E., Ghibelli, L., Nanoscale, 3
(2011) 1411–1420.
194. Colon, J., Herrera, L., Smith, J., Patil, S., Komanski, C., Kupelian, P.,
Seal, S., Jenkins, D.W., Baker, C.H., Nanomedicine, 5 (2009) 225–231.
195. Das, M., Patil, S., Bhargava, N., Kang, J.F., Riedel, L.M., Seal, S.,
Hickman, J.J., Biomaterials, 28 (2009) 1918–1925.
196. Colon, J., Hsieh, N., Ferguson, A., Kupelian, P., Seal, S., Jenkins, D.W.,
Baker, C.H., Nanomedicine, 6 (2010) 698–705.
197. Tarnuzzer, R.W., Colon, J., Patil, S., Seal, S., Nano Lett., 5 (2005)
2573–2577.
198. Chen, J.P., Patil, S., Seal, S., McGinnis, J.F., Nat. Nanotech., 1 (2006)
142–150.
199. Tang, B., Zhuo, L.H., Ge, J.C., Wang, G.L., Shi, Z.Q., Niu, J.Y., Chem.
Commun., 28 (2005) 3565–3567.
200. Korsvik, C., Patil, S., Seal, S., Self, W.T., Chem. Commun., 10 (2007)
1056–1058.

b1469_Ch-06.indd 357 4/8/2013 12:32:44 PM


b1469 Catalysis by Ceria and Related Materials

358 Z.-X. Li et al.

201. Perez, J.M., Asati, A., Nath, S., Kaittanis, C., Small, 4 (2008) 552–556.
202. Vincent, A., Babu, S., Heckert, E., Dowding, J., Hirst, S.M.,
Inerbaev, T.M., Self, W.T., Reilly, C.M., Masunov, A.E., Rahman, T.S.,
Seal, S., ACS Nano, 3 (2009) 1203–1211.
203. Asati, A., Santra, S., Kaittanis, C., Nath, S., Perez, J.M., Angew. Chem.,
Int. Ed., 48 (2009) 2308–2312.
204. Heckert, E.G., Karakoti, A.S., Seal S., Self, W.T., Biomaterials, 29 (2008)
2705–2709.
205. Pirmohamed, T., Dowding, J.M., Singh, S., Wasserman, B., Heckert, E.,
Karakoti, A.S., King, J.E.S., Seal,S., Self, W.T., Chem. Commun., 46 (2010)
2736–2738.
206. Mandoli, C., Pagliari, F., Pagliari, S., Forte, G., Di Nardo, P.,
Licoccia, S., Traversa, E., Adv. Funct. Mater., 20 (2010) 1617–1624.
207. Sobek, J.M., Talburt, D.E., J. Bacteriol., 1 (1968) 47–51.
208. Limbach, L.K., Li, Y.C., Grass, R.N., Brunner, T.J., Hintermann, M.A.,
Muller, M., Gunther, D., Stark, W.J., Environ. Sci. Technol., 39 (2005)
9370–9376.
209. Thill, A., Zeyons, O., Spalla, O., Chauvat, F., Rose, J., Auffan, M.,
Flank, A.M., Environ. Sci. Technol., 40 (2006) 6151–6156.
210. Xia, T., Kovochich, M., Liong, M., Madler, L., Gilbert, B., Shi, H.,
Yeh, J.I., Zink, J.I., Nel, A.E., ACS Nano, 2 (2008) 2121–2134.
211. Patsalas, P., Logothetidis, S., Metaxa, C., Appl. Phys. Lett. 81 (2002)
466–468.
212. Yablonovitch, E., Phys. Rev. Lett., 58 (1987) 2059–2062.
213. John, S., Phys. Rev. Lett., 58 (1987) 2486–2489.
214. Waterhouse, G.I.N., Metson, J.B., Idriss, H., Sun-Waterhouse, D.,
Chem. Mater., 20 (2008) 1183–1190.
215. Waterhouse, G.I.N., Waterland, M.R., Polyhedron, 26 (2007) 356–368.
216. Masui, T., Fujiwara, K., Machida, K.I., Adachi, G.Y., Sakata, T.,
Mori, H., Chem. Mater., 9 (1997) 2197–2204.
217. Li, R.X., Yabe, S., Yamashita, M., Momose, S., Yoshida, S., Yin, S.,
Sato, T., Solid State Ionics, 151 (2002) 235–241.
218. Van Leeuwen, R.A., Huang, C.J., Kammler, D.R., Switzer, J.A., J. Phys.
Chem., 99 (1995) 15247–15252.
219. Guo, S., Arwin, H., Jacobsen, S.N., Jarrendahl, K., Helmersson, U.,
J. Appl. Phys., 77 (1995) 5369.

b1469_Ch-06.indd 358 4/8/2013 12:32:44 PM


b1469 Catalysis by Ceria and Related Materials

2D and 3D Ceria-Based Nanoarchitectures 359

220. Orel, Z.C., Orel, B., Phys. Status Solid B, 186 (1994) K33–36.
221. Tsunekawa, S., Fukuda, T., J. Appl. Phys., 87 (2000) 1318.
222. Tsunekawa, S., Sahara, R., Kawazoe, Y., Kasuya, A., Mater. Trans., JIM,
41 (2000) 1104–1147.
223. Inoue, M., Kimura, M., Inui, T., Chem. Commun., 11 (1999) 957–958.
224. Yin, L.X., Wang, Y.Q., Pang, G.S., Koltypin, Y., Gadanken, A., J. Colloid
Interface Sci. 246 (2002) 78–84.
225. Wang, H., Zhu, J.J., Zhu, J.M., Liao, X.H., Xu, S., Ding, T., Chen, H.Y.,
Phys. Chem. Chem. Phys. 4 (2002) 3794–3799.
226. Si, R., Zhang, Y.W., You, L.P., Yan, C.H., Angew. Chem., Int. Ed., 44
(2005) 3256–3260.
227. Yabe, S., Sato, T., J. Solid State Chem., 171 (2003) 7–11.
228. El-Toni, A.M., Yin, S., Hayasaka, Y., Sato, T., J. Mater. Chem., 15 (2005)
1293–1297.
229. El-Toni, A.M., Yin, S., Sato, T., Appl. Surf. Sci., 252 (2006) 5063–5070.
230. Godinho, M., Ribeiro, C., Longo, E., Leite, E.R., Cryst. Growth & Des.,
8 (2008) 384–386.
231. Wang, Z.L., Quan, Z.W., Lin, J., Inorg. Chem., 46 (2007) 5237–5242.
232. Babu, S., Schulte, A., Seal, S., Appl. Phys. Lett., 92 (2008) 123112.
233. Yuan, Q., Duan, H.H., Li, L.L., Sun, L.D., Zhang, Y.W., Yan, C.H.,
J. Colloid Interface Sci., 335 (2009) 151–167.
234. Kaneko, K., Inoke, K., Freitag, B., Hungria, A.B., Midgley, P.A.,
Hansen, T.W., Zhang, J., Ohara, S., Adschiri, T., Nano Lett., 7 (2007)
421–425.

b1469_Ch-06.indd 359 4/8/2013 12:32:44 PM


b1469 Catalysis by Ceria and Related Materials

This page intentionally left blank

b1469_Ch-06.indd 360 4/8/2013 12:32:44 PM


b1469 Catalysis by Ceria and Related Materials

CHAPTER 7

CORE-SHELL-TYPE MATERIALS
BASED ON CERIA
Matteo Cargnello,1 Raymond J. Gorte,2
and Paolo Fornasiero1
1
Department of Chemical and Pharmaceutical Sciences,
University of Trieste, Via L. Giorgieri, 1, 34127 Trieste, Italy
2
Department of Chemical and Biomolecular Engineering,
University of Pennsylvania, 311A Towne Building,
220 South 33rd Street, Philadelphia, PA, 19104, USA

7.1 Introduction
In recent years, there have been many advances in the synthetic
methods for controlling the size and shape of materials at the
nanometer scale for applications in a number of fields, including
electronics, biology, materials science, catalysis, and medicinal
chemistry.1 The ability to tailor the structure of these materials to
obtain precise functionalities has been extensively highlighted in
the literature. In many of these applications, the interaction between
different components is a key factor for achieving the desired prop-
erties, so that the level of complexity in these materials is high.
A nice demonstration of this was given by Banin and co-workers,
who showed that the photo-induced charge separation between
CdS–PdO and CdS–Pd4S heterostructures, with applications for the
photo-reduction of water, could be obtained by tuning the morphol-
ogy of the material.2 Another interesting example is the synthesis of
binary nanocrystal superlattices using monodisperse nanoparticles

361

b1469_Ch-07.indd 361 4/8/2013 12:33:27 PM


b1469 Catalysis by Ceria and Related Materials

362 M. Cargnello, R. J. Gorte and P. Fornasiero

of two different materials using self-assembly methods.3 In this case,


the authors reported collective properties for the binary material
that were different from the single components.
Among the interesting nanostructures that can be prepared,
core-shell materials are especially intriguing because the close physi-
cal and electronic contact between the core and shell phases could
provide novel catalytic properties. This has been strongly pursued
for bimetallic catalysts, especially applications in fuel-cell elec-
trodes,4 where the metal core effectively modifies the electronic
properties of the atoms in the metal shell. Cooperative interactions
between the two components in a core-shell structure can also mani-
fest themselves in more complex ways. For example, the close
connection between the phases in core-shell materials allows rela-
tively easy modification following external stimuli. This was demon-
strated for Rh–Pd and Pt–Pd particles, which undergo changes in
morphology in response to the external environment, leading to
dynamic core-shell materials that could be used under either reduc-
ing or oxidizing conditions as catalysts with totally different
properties.5 In another example, large changes in the surface-
enhanced Raman scattering (SERS) signal intensity were observed
for dye molecules entrapped in Au–Au, core-shell nanostructures,
depending on the interior gap size.6
In this chapter, we will focus on core-shell structures in which
the core is a metal but the shell is ceria or some mixed oxide of
ceria. For catalytic applications, the oxide shell should probably be
considered at least somewhat porous, so that reactant molecules can
access sites on the metal core; however, there have been suggestions
that the presence of a group VIII metal in core-shell particles7 or
metal cations within the ceria8 can modify the properties of the
ceria. More generally, there are two primary reasons to consider
core-shell nanoparticles in which the shell is an oxide:

(i) The oxide coating can prevent the coalescence of the metal par-
ticles, especially when the reactions are to be carried out at
medium to high temperatures. At higher temperatures, metal
sintering is one of the main causes for the deactivation of

b1469_Ch-07.indd 362 4/8/2013 12:33:27 PM


b1469 Catalysis by Ceria and Related Materials

Core-Shell-Type Materials Based on Ceria 363

heterogeneous catalysts9,10 and the oxide shell would be expected


to prevent metal particles from coming together.
(ii) Reaction rates are enhanced by maximizing the number of
sites at the metal/ceria interface. Compared to precious metals
on other oxide supports, significantly enhanced rates have
been observed with ceria-supported metals for the water-gas-
shift,11 methane-steam-reforming,12,13 and CO2-reforming14,15
reactions. The role of ceria in these reactions is at least partially
associated with the ability of cerium to cycle between Ce+3 and
Ce+4, with the transfer of oxygen to and from the supported
metal playing an important role.16 For this kind of promotion,
close contact between the metal and the ceria support is critical
for enhanced reaction rates. It has also been suggested that
electron transfer between the oxide support and the metal
could modify the properties of the metal for small metal parti-
cles.17 Although long-range electron transfer from an insulat-
ing or semiconducting oxide, like ceria, to a metal is unlikely
to affect the properties of the larger metal particles,18 metal
atoms that are in direct contact with the oxide could be
affected, either by bonding interactions or by through-space
electric fields at the metal/oxide boundary. In both cases,
these interactions are relatively short range, so that the strong-
est interactions are expected to be within approximately 1 nm
of the metal/support interface.

Indeed, the ability to tune the structure and the interactions


between the oxide and metal phases is pivotal in a number of sup-
ported-metal catalysts, so that some supports are now recognized as
promoters or co-catalytic phases, rather than simple support materi-
als. The example of ceria-based water-gas shift (WGS) catalysts has
already been mentioned;11 also, recent reports suggest that the spe-
cial reactivity of Au–TiO2 materials also occurs through a dual-site
mechanism, implying the direct involvement of the support sites in
the activation of oxygen for CO and H2 oxidation reactions.19,20
Where more “inert” supports are used, the core-shell configuration
is still important for limiting sintering of the metal phase.21–23 The

b1469_Ch-07.indd 363 4/8/2013 12:33:27 PM


b1469 Catalysis by Ceria and Related Materials

364 M. Cargnello, R. J. Gorte and P. Fornasiero

combination of core-shell encapsulation and the use of reactive


supports in the shell has been demonstrated to both limit sintering
and enhance reactivity.24
The objective of this chapter is to provide a brief overview of
recent progress in the preparation of core-shell structures using
ceria as the shell. This approach, despite being relatively new, has
already been proven to provide materials with novel properties that
are different from those of previously known M/CeO2 materials.
Particular focus will be given to the unique behavior of these materi-
als under catalytic conditions, together with some characterization
issues that they pose.

7.2 Synthesis of Core-Shell Materials Based on Ceria


The preparation of core-shell structures, where a metal oxide is used
as the shell, generally relies on the synthesis of preformed metal
particles (or their precursors) and subsequent coating with the
metal oxide. In the case of ceria, the preparation of the core-shell
catalysts has been accomplished using one of three procedures, or
some combination of these. The general idea behind each of the
procedures is shown diagrammatically in Fig. 7.1 and can be sum-
marized as follows:

(a) The preformed metal particles (or their precursors) can be


precipitated together with the metal-oxide precursor.
(b) A microemulsion can be used as a nanoreactor, with the metal
particles first prepared by reduction and then coated with the
metal-oxide precursor to form the encapsulated metal.
(c) Direct modification and functionalization of the metal particle
surface allow sequential reaction with the metal-oxide precur-
sor and then produce a coating on the particles.

The detailed procedures required to produce other core-shell struc-


tures have been reviewed elsewhere and we will provide only a brief
overview of those involving ceria as the shell.10,25

b1469_Ch-07.indd 364 4/8/2013 12:33:27 PM


b1469 Catalysis by Ceria and Related Materials

Core-Shell-Type Materials Based on Ceria 365

Figure 7.1 Synthetic strategies to obtain metal@ceria core-shell structures.


(a) Co-precipitation of either preformed metal particles or metal particle precur-
sors and the metal-oxide precursor. (b) Microemulsion. (c) Direct functionaliza-
tion of preformed metal particles.

7.2.1 Co-precipitation
The co-precipitation method is based on the precipitation either of
a mixture of metal and metal-oxide precursors or of preformed
metal particles and the metal-oxide precursor. Although precipita-
tion is a common phenomenon, the details behind the steps needed
to understand how core-shell materials are formed requires recog-
nition of the fact that precipitation goes through a number of
different stages, including nucleation, Ostwald ripening of the crys-
tallites, and aggregation processes.25 The preparation of core-shell
structures with this method requires good control over the

b1469_Ch-07.indd 365 4/8/2013 12:33:27 PM


b1469 Catalysis by Ceria and Related Materials

366 M. Cargnello, R. J. Gorte and P. Fornasiero

experimental conditions in order to maximize the interaction


between the metal and the metal-oxide precursor before their pre-
cipitation. Precipitation is usually a result of a chemical reaction,
with the preferred methods requiring the use of weak or strong
bases as precipitating agents. Therefore, many co-precipitation syn-
theses use aqueous solutions or suspensions of the precursors.
In 2008, Zhang and co-workers prepared “Ir-in-ceria” catalysts
for the preferential oxidation of CO (the PROX reaction) using a
co-precipitation method.26 Ir and Ce salts were co-precipitated by
addition to a warm solution of NaOH. Along with precipitation, the
basic conditions and the slightly elevated temperature favored the
reduction of Ir(IV) to Ir(0) through the oxidation of Ce(III) to
Ce(IV). The effective encapsulation of metallic Ir particles was sug-
gested by in situ diffuse reflectance infrared Fourier-transform spec-
troscopy (DRIFTS), which showed a very weak CO chemisorption
signal when compared to Ir/CeO2 catalysts prepared by conven-
tional impregnation procedures.
Some of us recently reported the preparation of Au@CeO2 cata-
lysts for the PROX reaction.27 The samples were prepared by a modi-
fied co-precipitation technique. Preformed, 1.7 nm Au nanoparticles
were synthesized using a mixture of two thiols as protecting agents:
the first thiol included a polyoxethylene chain to impart solubility in
water and a hydrocarbon chain close to the gold surface to provide
a tightly packed protective shield for the nanoparticle core; the sec-
ond thiol, 16-mercaptohexadecanoic acid, was included to ensure
the presence of carboxyl groups at the external surface of the parti-
cles. These latter groups were then exploited to form complexes
with Ce(IV) cations and to direct the co-precipitation of Ce(OH)x
species around the preformed Au nanoparticles. This was accom-
plished by heating an aqueous solution of Au particles and
Ce(NH4)2(NO3)6 in urea. Urea, under heating, slowly hydrolyzed to
CO2 and ammonia, raising the pH and causing a slow precipitation
of Ce(OH)x. The carboxyl groups on the Au particles ensured that
a Ce matrix was formed around the nanoparticles. The presence of
the second amphiphilic thiol on the Au monolayer allowed the
porous structure to be tuned during the synthesis. This method was

b1469_Ch-07.indd 366 4/8/2013 12:33:28 PM


b1469 Catalysis by Ceria and Related Materials

Core-Shell-Type Materials Based on Ceria 367

shown to be particularly flexible in that Au particles with different


average diameters could be used and the Au loading could be
easily tuned.
Shinjoh and colleagues used similar concepts to prepare a
CeO2–Ag catalyst that had a “rice-ball” configuration, where the
shell was composed of a layer of small CeO2 crystallites.28 Again,
Ce(III) and Ag(I) nitrates were co-precipitated by instantaneous
addition to an aqueous ammonia solution and then subjected to
a hydrothermal treatment. The final structure of the CeO2–Ag
nanocomposites was that of spherical agglomerates, roughly
100 nm in diameter (Fig. 7.2(a)). The authors obtained a thin
slice of a single rice-ball using a focused ion-beam technique
(Fig. 7.2(b)); using transmission electron microscopy (TEM),
they demonstrated that Ag effectively formed a core 28 nm in
diameter (in accordance with X-ray diffraction (XRD) data),
surrounded by 14-nm CeO2 crystallites. The TEM images clearly
demonstrated that the CeO2 shell produced by the ceria crystal-
lites was porous.
An interesting observation about the CeO2–Ag structures is that
the resulting materials had a 1:1 Ce/Ag molar ratio, regardless of
the initial excess of Ag used during synthesis. This was rationalized
by invoking a redox mechanism in which [Ag(NH3)2]+ species
formed by the excess ammonia during co-precipitation were
reduced to Ag(0) by the oxidation of Ce(OH)3 to Ce(OH)4, which
in turn was dehydrated to form CeO2. This process has the drawback
that the Ag and CeO2 content in the catalysts cannot be controlled
and the final Ag loading (~40 wt%) is high and cannot be tuned.
Our groups also used a co-precipitation technique to prepare
Pd@CeO2 catalysts, starting from preformed Pd particles that were
prepared using polyvinylpyrrolidone (PVP) as protecting agent.29
The preparation of the 1.5 nm, protected-Pd nanoparticles was
based on the reduction of PdCl2 by means of a warm solution of
methanol in the presence of PVP.30 Subsequently, an ammonia solu-
tion was added to an aqueous solution containing both the PVP–Pd
particles and Ce(NO3)3, causing precipitation of CeO2 and embed-
ding of the Pd particles. Pd/CeO2 catalysts prepared by classical

b1469_Ch-07.indd 367 4/8/2013 12:33:28 PM


b1469 Catalysis by Ceria and Related Materials

368 M. Cargnello, R. J. Gorte and P. Fornasiero

Figure 7.2 (a) SEM of Ag@CeO2 rice-ball structures formed by co-precipitation.


(b) TEM cross section of a single structure showing the Ag core and the ceria shell
formed of multiple small crystallites. (c), (d) Schematic models of the structures.
Reprinted from Kayama et al.28 Copyright 2010 the American Chemical Society.

impregnation of a ceria prepared by the same precipitation methods


showed different reduction behavior in temperature-programmed
reduction (TPR) experiments with respect to the corresponding
embedded materials. The TPR results indirectly suggested better
contact between the metal and the ceria in the embedded sample.
Conventional high-resolution transmission electron microscopy
(HRTEM) investigations of the embedded catalyst were unable to
detect exposed Pd particles dispersed on the support, further indi-
cating that the metal particles were effectively within the ceria.

b1469_Ch-07.indd 368 4/8/2013 12:33:28 PM


b1469 Catalysis by Ceria and Related Materials

Core-Shell-Type Materials Based on Ceria 369

Using a similar approach, Xu and co-workers also started from


preformed metal particles to prepare Pd and Pt@CeO2 samples.31,32
In their case, the preformed Pt and Pd colloids were protected by
citrate and PVP, respectively, mixed with CeCl3 in an aqueous solu-
tion with urea and subjected to a hydrothermal treatment. The
hydrolysis of urea in warm water formed ammonia, turning the solu-
tion basic and precipitating Ce(OH)x species around the preformed
particles. The final material was then composed of Pd and Pt cores
embedded in a ceria matrix (Fig. 7.3).
There were remarkable differences between the Pd and Pt sam-
ples in the work of Xu et al.31,32 With the Pd@CeO2 samples, small
(2–4 nm) particles were homogeneously dispersed inside the oxide
spheres that formed during precipitation and the initial Pd particle
size was essentially preserved in the final calcined structures. By con-
trast, the Pt@CeO2 samples showed few and very large (100 nm) Pt
cores embedded inside the ceria particles as a result of the agglom-
eration of metal particles during the hydrothermal and calcination
processes. This was due to the different protecting agent used for
the synthesis of the particles (PVP versus citrate), with PVP demon-
strating itself to be better at preventing the undesired agglomeration
of the particles during hydrothermal treatment. It is also noteworthy
that the precipitation of ceria by these procedures, in the absence of
the metal particles, produced only irregularly-shaped ceria struc-
tures, indicating the templating effect of the metal particles during
the precipitation of Ce(OH)x. TEM (Fig. 7.3), energy dispersive
X-ray spectroscopy (EDS), and X-ray photoelectron spectroscopy
(XPS) analysis confirmed the core-shell nature of the composites.

7.2.2 Microemulsion-based synthesis


Microemulsions are thermodynamically stable systems in which
micelles are formed by surfactant molecules used to disperse two
immiscible liquids, usually water and a hydrocarbon solvent. Two
configurations can be obtained depending on the relative ratio of
the amounts in each solvent phase, so that both oil-in-water and
water-in-oil (also called reverse) microemulsions can be formed.

b1469_Ch-07.indd 369 4/8/2013 12:33:28 PM


b1469 Catalysis by Ceria and Related Materials

370 M. Cargnello, R. J. Gorte and P. Fornasiero

Figure 7.3 Pd@CeO2 structures prepared by precipitation of preformed PVP-


protected Pd particles and CeCl3 under basic conditions. (a) SEM image of the cal-
cined structures together with (b) a size distribution, (c) TEM image, (d) HRTEM
image showing the presence of Pd particles embedded inside the ceria matrix.
(e), (f) TEM images of the starting PVP-protected Pd colloids. Reprinted from
Zhang et al.32 Copyright 2011 the American Chemical Society.

b1469_Ch-07.indd 370 4/8/2013 12:33:28 PM


b1469 Catalysis by Ceria and Related Materials

Core-Shell-Type Materials Based on Ceria 371

Often, the presence of a co-surfactant (such as short chain alcohols


or amines) may help form the microemulsion by reducing the repul-
sion between the polar heads of surfactant molecules in the micelle
layer. These systems are well known and widely employed in the
preparation of nanocrystalline materials.33
The method used by the group of Tsang et al. to prepare Pt@CeO2
materials and other combinations of mono- and bimetallic particles
inside a ceria shell34,35 was based on an inverse microemulsion, pre-
pared using cetyltrimethylammonium bromide (CTAB) as sur-
factant in toluene and dissolving Pt(II) and Ce(III) salts in the water
phase. An aqueous solution of NaOH was added to the microemul-
sion in order to precipitate both precursors inside the micelles, after
which the mixture was aged for six days. By first precipitating Pt(II)
with NaOH and adding the Ce(III) salt later, the authors reported
that metal encapsulation was improved, e.g. the metal particles were
totally encapsulated inside the ceria shell, as demonstrated by chem-
isorption experiments. The authors claimed that the oxidation of
Ce(III) to Ce(IV) was coupled to the reduction of Pt(II) to Pt(0)
and to the formation of metallic particles in the core. XPS data sup-
ported this picture, since a very low degree of Ce(III) oxidation was
observed in the absence of Pt.35 It was suggested that, after forming,
the metal particles were then encapsulated inside CeO2 during the
aging process by polymerization and condensation of Ce hydroxyl
precursor species, so that the aging step was critical for the proce-
dure (Fig. 7.4).
Further proof that the ceria-embedded metals were formed in
the micelles was that the ceria crystallite size progressively decreased
with increasing Pt metal content from 1 to 5 wt% as a result of the
interaction of Pt and Ce precursors inside the micelles. This was due
to the increase in the size of the metal core precursor inside the
micelles, which in turn limited the growth of the ceria shell.
In our laboratories, a different strategy was used to prepare Pd@
CeO2 catalysts from microemulsions.29 The microemulsion was first
prepared using CTAB as surfactant and 1-butanol as co-surfactant in
isooctane and water, with Pd(NO3)2 employed as Pd precursor and
dissolved in the aqueous phase. Hydrazine (N2H4) was directly

b1469_Ch-07.indd 371 4/8/2013 12:33:28 PM


b1469 Catalysis by Ceria and Related Materials

372 M. Cargnello, R. J. Gorte and P. Fornasiero

Figure 7.4 TEM micrograph of Pt(5 wt%)@CeO2 samples prepared by microe-


mulsion showing ceria crystallites (∼4 nm); the inset shows the enlarged image of
an isolated nanoparticle with Pt enrichment in the core area. Reprinted from
Yeung et al.34 Copyright 2005 the American Chemical Society.

added to the microemulsion to reduce Pd(II) to Pd(0). An appro-


priate amount of a cerium(IV) alkoxide dissolved in isooctane, a
compound that is very sensitive to hydrolysis, was added, which in
turn hydrolyzed with the water present in the micelles of the micro-
emulsion, thus causing the precipitation of Ce(OH)4 species around
the preformed Pd particles.
In another very recent report, a microemulsion-based method
was used by Kaneda and co-workers to prepare Ag@CeO2 compos-
ites.36 The synthetic procedure coupled several of the strategies dis-
cussed earlier in this chapter; it was based on a redox reaction
between Ag(I) and Ce(OH)3 species co-precipitated in water/
ammonia pools of the microemulsion. The result was very similar to
the Ag@CeO2 “rice-balls” described in Section 7.2.1,28 but the
dimensions were much smaller for both the Ag-rich core (~10 nm)
and the CeO2 shell (~8 nm) formed by 3–5 nm ceria crystallites.
Again, the ceria shell was found to be porous, a fundamentally
important property for catalytic applications (see Section 7.3).
Though the Ag loading was high (40 wt%), similar to the Ag@CeO2
“rice-balls,” the Ag–CeO2 interfacial surface area of these assemblies
was much larger due to the reduced dimensions.

b1469_Ch-07.indd 372 4/8/2013 12:33:28 PM


b1469 Catalysis by Ceria and Related Materials

Core-Shell-Type Materials Based on Ceria 373

7.2.3 Other methods


Another method for preparing core-shell nanoparticles involves the
direct modification and functionalization of the metal particle
surface to allow sequential reaction with the metal-oxide precursor.
This approach has been used frequently in the preparation of core-
shell structures with SiO2 as the shell, primarily because the
chemistry of Si is very similar to that of carbon, so that typical
organic reactions can then be applied. Silicon alkoxides are well-
known precursors. Many of the silicon alkoxides are commercially
available and easy to functionalize. For example, this advantage was
employed by Liz-Marzán and co-workers in their pioneering work on
silica coating of Au nanoparticles.37 The surface of the Au particles
was first made “vitrophilic” by protection with a silane bearing an
amine functionality, (3-aminopropyl)trimethoxysilane (APS). This
silane was able to bind the Au core with the amine group while
exposing a silane as an anchoring point to start growing the SiO2
layer through a classical Stöber procedure. In this way, both the
diameter of the Au core and the thickness of the silica layer were
tunable, depending on the size of the initial Au particles used and
the number of silica precursors added during the synthesis.
This first initial report stimulated much interest in this type of
chemistry. In 2007, Ying and co-workers reported the synthesis of
core-shell structures that were dispersible in water through the use
of several functionalized silanes.38 The key feature of this work was
that the particles were directly protected with functionalized silanes
(Fig. 7.5). The most important contribution of this approach is that
different types of particles (oxides, metals, semiconductors) can be
at the center of these core-shell structures.
Because alkoxides of less electronegative metals are more sus-
ceptible to hydrolysis and therefore more difficult to functionalize,
we chose to follow a different approach to obtain Pd@ceria core-
shell structures. A schematic showing the basic procedure is given in
Fig. 7.6.39
In the first step, we synthesized Pd nanoparticles protected
by 11-mercaptoundecanoic acid (MUA), referred to here as

b1469_Ch-07.indd 373 4/8/2013 12:33:28 PM


b1469 Catalysis by Ceria and Related Materials

374 M. Cargnello, R. J. Gorte and P. Fornasiero

Figure 7.5 Silane conjugation chemistry used to cover several oxide and non-
oxide particles with a thin layer of silica. Reprinted from Jana et al.38 Copyright 2007
the American Chemical Society.

Figure 7.6 Procedure to obtain dispersible Pd@CeO2 core-shell nanostructures by


self-assembly of a cerium(IV) alkoxide around preformed Pd particles protected by
11-mercaptoundecanoic acid. Reprinted from Cargnello et al.39 Copyright 2010 the
American Chemical Society.

b1469_Ch-07.indd 374 4/8/2013 12:33:28 PM


b1469 Catalysis by Ceria and Related Materials

Core-Shell-Type Materials Based on Ceria 375

MUA–Pd particles, through an optimized simple, single-phase,


single-step process.40 The thiol moiety of the MUA molecule pro-
tected the metal core, while the carboxyl group was available at
the surface of the particles for further reaction. The carboxyl
groups were used to coordinate Ce(IV) alkoxide41 due to the bet-
ter coordinating ability of carboxylic acids with respect to alkox-
ides. The presence of dodecanoic acid partially protected the
remaining alkoxide tails of the Ce(IV) centers, so that, after
hydrolysis, the Pd@CeO2 structures that were produced still had
the hydrophobic tails of the dodecanoic acid molecules at the
surface, guaranteeing the particles could be dispersed in low
polar solvents. The demonstration of the core-shell nature of the
assemblies was obtained through a combination of atomic force
microscopy (AFM), TEM, and elemental mapping by electron
energy loss spectroscopy (EELS). Results from the EELS mapping
are shown in Fig. 7.7. Figures 7.7(A1) and (A2) show images of
two different regions of the sample obtained using the Ce post-
edge, where all the major elements (Ce, Pd, and S) contributed
to the final micrographs. The contributions from lighter ele-
ments, such as C and O, were limited due to their very low scatter-
ing factors. Removing the pre-edge contribution (mainly Pd and S)
from the total post-edge signals (Ce, Pd, and S) gave maps of the
Ce distribution (Figs. 7.7(B1) and (B2)), which were very similar
to Figs. 7.7(A1) and (A2). Collecting the signals from the Ce pre-
edge, which had the Ce contribution removed from the image,
resulted in Figs. 7.7(C1) and (C2). Individual particles having Pd
and S were clearly visible, corresponding to the metallic cores.
This was consistent with a Ce-containing layer surrounding the Pd
particles and demonstrated the core-shell nature of the
nanostructures.
Current core-shell structures have the advantage of being dis-
persible in organic solvents, without any sign of agglomeration after
several months. Therefore, they could be easily adsorbed onto sup-
ports and used for different applications, as we will demonstrate
later in this chapter.

b1469_Ch-07.indd 375 4/8/2013 12:33:29 PM


b1469 Catalysis by Ceria and Related Materials

376 M. Cargnello, R. J. Gorte and P. Fornasiero

Figure 7.7 EELS maps of two different regions (distinguished as 1 and 2) of


dispersible Pd@CeO2 nanostructures. (A) Ce post-edge, all of the elements
contribute to the image. (B) Ce map, only Ce contributes. (C) Ce pre-edge, only
Pd and S contribute. The scale bars represent 20 nm. Reprinted from Cargnello
et al.39 Copyright 2010 the American Chemical Society.

7.3 Applications
7.3.1 Water gas shift reaction (WGSR)
The WGSR, shown in Eq. (7.1), is a key step in a number of indus-
trial applications:

CO + H2O = CO2 + H2, ∆H0 = −40.6 kJ.mol−1 (7.1)

The reaction is used to control the ratios of CO, CO2, and H2 in


synthesis gas and it is directly or indirectly relevant to several impor-
tant industrial catalytic processes, including methanol synthesis,42,43
ammonia synthesis, Fisher–Tropsch synthesis, coal gasification, and
catalytic combustion. Moreover, it is involved in all reforming
processes starting from hydrocarbons44 or higher alcohols (such as

b1469_Ch-07.indd 376 4/8/2013 12:33:29 PM


b1469 Catalysis by Ceria and Related Materials

Core-Shell-Type Materials Based on Ceria 377

ethanol45) or methanol.46 The importance of the WGSR has also


been demonstrated in the aqueous phase reforming of oxygenates.47
Current, commercially available WGSR catalysts have important
drawbacks, in that they can be toxic, pyrophoric, and prone to sin-
tering if not handled very carefully.48 Alternative catalysts are there-
fore needed, especially for fuel processing in fuel-cell applications,
where the presence of a plant operator is impractical. Various mate-
rials have been proposed, ranging from catalysts based on coinage
metals, such as Ni and Cu, to ones based on noble metals, including
Au, Pd, and Pt.49,50 Since the demonstration that ceria-supported
precious metals can exhibit rates approaching that of low-tempera-
ture, WGS catalysts,11 ceria-based catalysts have received a great deal
of attention.
Regarding the use of core-shell catalysts, Tsang and co-workers
were among the first to study the activity of Pt@CeO2 and related
systems for the WGSR.7,34,35,51,52,53 They reported that the best cata-
lytic activities were obtained following sequential addition of Pt and
then Ce to the microemulsion, with the use of NaOH as base. They
also indicated that an aging time of six days was important.52 A major
finding in their work was that methanation activity (hydrogenation
of CO and CO2 with H2 to give methane) of the embedded materials
was absent, or at least much lower than that found on conventional
Pt/CeO2 catalysts with exposed metal sites (see below). Despite this,
these authors reported the WGS activity to be similar to that of
Cu/ZnO and conventional Pt/CeO2 samples, thus implying that the
active sites for WGS are very different from those needed for
methanation.53
However, some of the claims from their work remain controver-
sial. In their view, the sites responsible for WGS activity on their
Pt@ceria catalysts are exclusively located on the ceria surface, with
the embedded metal essentially promoting the activity of the ceria
by modifying the electronic states of the oxide, even though it is
difficult to envision how these sites could carry out the reaction.
They based their conclusion on the following observations. First,
the absorption edge (O2p – Ce4f) of the ceria shell, measured using
diffuse ultraviolet reflectance, was shifted towards the blue in the

b1469_Ch-07.indd 377 4/8/2013 12:33:29 PM


b1469 Catalysis by Ceria and Related Materials

378 M. Cargnello, R. J. Gorte and P. Fornasiero

embedded systems compared to that observed with conventional


ceria. The authors attributed this to electronic interactions and
charge-transfer processes between the metals and the ceria. The
largest shift was observed for bimetallic Pt–Au@CeO2 systems, and
these were reported to give the highest WGS activities observed for
the core-shell series that was investigated.34 Second, TPR experi-
ments indicated that ceria reduction occurred at higher tempera-
tures on the embedded systems when compared to conventional
catalysts. This difference was believed to be due to the absence of
exposed Pt surfaces on the embedded samples, with the result that
no extensive H2 spill-over occurred during TPR. Third, in situ
Fourier transform infrared (FTIR) experiments exhibited negligi-
ble absorption in the ν(CO) stretching region associated with
CO adsorbed on Pt under reverse WGS conditions in the embed-
ded system, even though the sample showed good activity.7 Chem-
isorption experiments also showed unusually low CO uptakes for
the embedded systems, so that the calculated Pt particle size would
be 50 nm, even though XRD experiments were unable to identify
any peak due to such large crystallites. Finally, the rate law for the
WGSR exhibited a strong positive dependence for the reaction rate
on the concentration of water and a large negative dependence on
CO2.35 This led the authors to suggest that neither of the two
reaction mechanisms postulated for WGS on Pt ceria, namely the
redox mechanism11 or the bidentate formate mechanism,54,55 were
primarily responsible for reactions over Pt@ceria embedded
catalysts.
The WGSR was also used by our groups to test the thermal sta-
bility of Pd@CeO2 catalysts prepared by co-precipitation and micro-
emulsion.29 It has been shown that metal sintering is one of the
main causes of deactivation for the WGSR on conventional, ceria-
supported, noble-metal catalysts.9,56 The two embedded systems
(Pd@CeO2-CP and Pd@CeO2-ME, where CP and ME stand for
co-precipitation and microemulsion, respectively) were compared
with a conventional Pd/CeO2 catalyst prepared by incipient-wetness
impregnation (IMP) with the same Pd loading (1 wt%). The results
are shown in Fig. 7.8.

b1469_Ch-07.indd 378 4/8/2013 12:33:29 PM


b1469 Catalysis by Ceria and Related Materials

Core-Shell-Type Materials Based on Ceria 379

Figure 7.8 CO conversions at 250 °C (a), (b) and 300°C (c) under WGSR condi-
tions for (a) Pd/CeO2-IMP, (b) Pd@CeO2-CP, and (c) Pd@CeO2-ME catalysts after
different subsequent treatments. Reproduced from Cargnello et al.29 by permission
of the Royal Society of Chemistry.

b1469_Ch-07.indd 379 4/8/2013 12:33:29 PM


b1469 Catalysis by Ceria and Related Materials

380 M. Cargnello, R. J. Gorte and P. Fornasiero

The CP and IMP catalysts initially exhibited similar CO conver-


sions under WGS conditions at 250 °C (Fig. 7.8(a) and (b)), ~ 10%.
However, while a subsequent aging treatment at 400°C for 3 h
caused the conversion to drop to roughly 7% on both, the CP cata-
lyst showed a somewhat higher stability thereafter as observed by the
fact that the CO conversion of the IMP catalyst dropped even under
steady state conditions at 250 °C. Temperature-programmed oxida-
tion (TPO) treatments were performed in order to understand the
nature of the deactivation. A mild oxidative treatment at 400 °C
resulted in no appreciable recovery of the activity. This fact was con-
sistent with an irreversible deactivation of the catalysts caused by
sintering after the high-temperature aging treatment. A further,
stronger oxidative treatment at 600 °C was performed, which
resulted in the recovery of the initial catalytic activity. However, this
restoration of activity could be explained by the oxidative redisper-
sion of PdO after high-temperature treatments under oxygen.57–59
The minor enhancement in stability observed in the case of the
Pd@CeO2-CP catalyst was an indication that a thin ceria layer could
be effective for protecting the metal particles against sintering. The
activity of the Pd@CeO2-ME catalyst was therefore tested
(Fig. 7.8(c)). This sample initially showed lower CO conversion at
300 °C compared to the other samples, ~ 7%, but it exhibited very
good thermal stability, and no deactivation was observed either
after the first aging treatment at 400 °C for 3 h or after the same
treatment for 12 h. As a further demonstration that no sintering
processes occurred, TPO at 600 °C did not produce any increase in
catalytic activity, since no Pd redispersion was required for this
sample. The lower CO conversion over this sample was attributed
to the poor textural properties of the support. Indeed, during the
ME procedure, the ceria species condensed too early, producing a
very dense ceria layer with rather low surface area (18 m2.g−1 com-
pared to 80 m2.g−1 for the CP sample). Despite this, the embedding
of the metal phase inside this dense ceria layer provided an appro-
priate barrier to avoid metal sintering and consequently the catalyst
did not undergo deactivation under WGS conditions even at high
temperatures.

b1469_Ch-07.indd 380 4/8/2013 12:33:30 PM


b1469 Catalysis by Ceria and Related Materials

Core-Shell-Type Materials Based on Ceria 381

WGS was also the target reaction to test the activity and stability
of Pd@CeO2 core-shell structures prepared by self-assembly.39,60 As dis-
cussed in Section 7.2.3, these structures are dispersible in organic
solvents and can be adsorbed on relatively high-surface-area alu-
mina (100 m2·g−1) in order to isolate single entities on the surface of
the support. The samples that were studied are referred to here by
their composition (e.g. Pd(1%)@CeO2(9%)/Al2O3). The initial rate
for the Pd(1%)@CeO2(9%)/Al2O3 catalyst was very good, similar to
that observed over a conventional Pd(1%)/CeO2 sample. However,
the rate over the core-shell catalyst decreased dramatically over a
period of 1 hour, with the initial CO conversion at 350 °C going
from 12% to only 4% after 60 minutes. The loss of activity could be
completely reversed following a brief oxidation treatment in air at
the same temperature for 10 minutes, suggesting that the deactiva-
tion of the catalyst was not related to changes in metal particle size
but more likely to changes in the redox states of the components.
We found that the WGS rates of the embedded sample were lower
when it was subjected to a reduction pretreatment in H2 before col-
lecting the WGS data. All these observations pointed to a special and
unusual redox behavior of the ceria component of the catalyst
under WGS conditions.
To understand these characteristics we measured the oxidation
states of the Pd@CeO2/Al2O3 sample after various pretreatments
using oxygen titration experiments, with the results shown in
Table 7.1. The pretreatment in hydrogen caused a significant

Table 7.1 Oxygen uptake by the Pd@CeO2/Al2O3 catalyst following


reduction by pure H2 gas at 350 °C, oxidation in air at 350 °C fol-
lowed by exposure to WGS conditions at 350 °C, and reduction in H2
with subsequent partial oxidation by 25 torr H2O at 350 °C.
Oxygen uptake
Pretreatment (µmol O g−1 catalyst) O:Ce ratio
H2 200 1.61
Air, followed by WGS 150 1.70
H2, followed by steam 180 1.66

b1469_Ch-07.indd 381 4/8/2013 12:33:30 PM


b1469 Catalysis by Ceria and Related Materials

382 M. Cargnello, R. J. Gorte and P. Fornasiero

reduction of the catalyst, so that the O:Ce ratio in the catalyst was
calculated to be approximately 1.61. (This calculation did not take
into account the reduction of PdO to Pd because it was difficult to
distinguish between these two contributions.) This degree of reduc-
tion was similar to that observed for conventional Pd/CeO2 catalysts
after treatment in H2.77 More interesting are the other data in
Table 7.1. Pretreatment of the sample in air, followed by WGS, gave
an approximate O:Ce ratio of 1.70, which implies that the catalyst
existed in a reduced state under WGS conditions. The WGS environ-
ment is relatively oxidizing for ceria, so that a traditional catalyst
remains essentially oxidized following this pretreatment. Because Pd
will be reduced under WGS conditions, the calculation for the O:Ce
ratio should not be affected by the Pd in this case. The interesting
fact here was that the catalyst was effectively reduced by H2 or CO
but was not reoxidized by H2O and CO2. To further test this idea, we
prereduced the catalyst in dry H2 and then treated it under steam.
The O:Ce ratio was again approximately 1.66, clearly demonstrating
that the catalyst could not be significantly reoxidized by steam, in
sharp contrast with a conventional Pd/CeO2 system where steam is
a very effective oxidizing agent. All these observations suggested that
the ceria shell in the embedded system had very different properties
to bulk ceria, being much more easily reduced.
Although we did not provide a definitive explanation for this
behavior, we suggest that the Pd particles might act as template for
the growth of a disordered ceria shell, with a local structure that is
different from the usual bulk, fluorite structure. The fluorite struc-
ture is key for stabilizing Ce(IV),61 while Ce(III) is the most stable
species in many other environments. For example, Ce exists as
Ce(III) in CeVO4 no matter how oxidizing the environment.62 If a
significant fraction of the Ce atoms exist in a disordered structure
within the shell, Ce(III) could be stabilized after reduction. In this
regard, it is noteworthy that ceria powders can have very different
thermodynamic, redox properties when they are prepared at lower
temperatures.63,64
Another interesting feature of these core-shell catalysts is the
effect of reduction on the accessibility of reactants to the Pd.

b1469_Ch-07.indd 382 4/8/2013 12:33:30 PM


b1469 Catalysis by Ceria and Related Materials

Core-Shell-Type Materials Based on Ceria 383

Figure 7.9 Changes in shell morphology for oxidized and reduced Pd@CeO2/
Al2O3 core-shell catalysts. Reprinted from Wieder et al.60 Copyright 2011 the
American Chemical Society.

Combined diffuse-reflectance FTIR and CO chemisorption studies


showed that the embedded sample adsorbed CO when the ceria was
in an oxidized state but not after it was reduced. These effects are
analogous to what has been reported with titania- and ceria-sup-
ported catalysts following high-temperature reduction, where evi-
dence has been given for the oxide forming a film on the metal,17,65
although the reduction conditions in our present experiments were
much milder. However, we again propose that a geometric covering
by the ceria shell physically limits CO adsorption in the reduced
catalyst as shown schematically in Fig. 7.9.
In this picture, the oxidized CeO2 forms a shell with many cracks
and fissures that allow gases to adsorb on the Pd cores. Following
reduction, these fissures are closed due to either the decreased den-
sity of Ce2O3 compared to CeO2 or to stronger interactions between
reduced ceria and Pd that cause the spreading of the support onto
the particles.
In conclusion, despite some attractive properties for these
embedded Pd@CeO2/Al2O3 catalysts, they are not ideal for WGS.
Still, there are clear indications that core-shell-type materials have
very different properties from conventional catalysts. The easily
reducible, ceria shell in this new type of material could allow these
catalysts to find applications in other areas. Indeed, we will show
later in this chapter how these materials exhibit superior stability in
solid oxide fuel cell (SOFC) anodes (see Section 7.3.3).

b1469_Ch-07.indd 383 4/8/2013 12:33:30 PM


b1469 Catalysis by Ceria and Related Materials

384 M. Cargnello, R. J. Gorte and P. Fornasiero

7.3.2 Gas-phase oxidation reactions


CO oxidation in the presence of hydrogen, the so-called PROX reac-
tion, has received significant attention in recent years because of its
importance for the development of proton exchange membrane
(PEM) fuel cell systems for automotive applications.66 Current fuel-
cell materials are poisoned by CO impurities in common H2 feeds,
requiring the removal of CO, potentially by using the PROX reac-
tion.67 Regarding core-shell catalysts, Zhang and colleagues investi-
gated the activity of Ir@ceria catalysts, prepared by co-precipitation
methods, for the PROX reaction in a series of papers.26,68,69 Initial
tests under simple mixtures of CO, H2, and O2 demonstrated that
the Ir@ceria systems showed improved CO conversion and selectiv-
ity for CO2 in a wider temperature window compared to a conven-
tional Ir/ceria catalyst.26 The addition of H2O to the stream did not
change the performance of the embedded catalyst but water was
found to be fundamentally important for achieving high PROX
activity. Unfortunately, these authors observed a dramatic decrease
in activity when CO2 was introduced to the reactant mixture.
The preparation conditions for the catalysts in these studies
affected their performance. A study of precipitation conditions and
metal loading determined that the best catalysts were prepared with
strong precipitating agents (such as NaOH) and at intermediate Ir
loadings of 1.6 wt%. These observations suggested that the active
sites were formed by encapsulation of the Ir during the co-precipita-
tion procedure, with Ir(IV) being reduced to Ir(0) and simultane-
ously embedded inside the ceria matrix. The authors argued that
there must be a balance between the synthesis parameters in order
to obtain catalysts that activate CO, which is required for reactions
to occur, but do not over-activate H2, which would result in low selec-
tivity. A comparison of samples prepared in a similar way, but using
Au or Pt in place of Ir, showed that all the embedded systems were
good PROX catalysts, although each of the metals had a different
temperature window. The authors concluded that Ir was not directly
involved in the catalytic cycle but that its role was to activate surface
oxygen of ceria, similar to the model reported by Tsang and

b1469_Ch-07.indd 384 4/8/2013 12:33:30 PM


b1469 Catalysis by Ceria and Related Materials

Core-Shell-Type Materials Based on Ceria 385

colleagues for Pt@CeO2 systems (see Section 7.3.1). Based on the


observation that Ir@ceria samples were more active than Ir/ceria at
lower temperatures, the authors prepared a dual bed system with
Ir@ceria placed before an Ir/ceria catalyst, in order to widen the
operating-temperature window.68
The PROX reaction was also studied on Au@CeO2 catalysts.27
Embedded samples with 1 and 3 wt% Au loadings were compared to
catalysts prepared by a more traditional, deposition-precipitation-
with-urea (DPU) method following a procedure from Zanella et al.70
The DPU procedure is known to produce active Au catalysts for CO
oxidation. The aim of the PROX work was to assess the relative ther-
mal stability of the different formulations following simulated aging
procedures under realistic PROX conditions (1.0 vol% CO, 1.0 vol%
O2, 47.5 vol% H2, 17.5 vol% CO2, and 5.0 vol% H2O) at 150 °C for
16 h and at 250 °C for 20 h. The catalytic tests demonstrated that all
of the samples that were tested suffered deactivation after the aging
treatments due to the formation of carbonate species that blocked
the access to active Au sites. However, unlike the DPU catalysts, the
embedded samples showed a complete recovery of the initial activity
following TPO. After TPO, the DPU catalysts were only partially
restored to their initial activity. The loss of activity in the DPU cata-
lysts was due to sintering of the Au, which caused coarsening of the
metal particles even at 250 °C, as demonstrated by high-angle annu-
lar dark-field scanning transmission electron microscopy (HAADF-
STEM) investigations. By contrast, the encapsulation of Au particles
inside the porous CeO2 matrix in the embedded samples prevented
this unwanted phenomenon and therefore contributed to the stabil-
ity of these systems under reaction conditions.
Core-shell catalysts have also been studied for catalytic soot com-
bustion, a gas-phase oxidation reaction used for the removal of
diesel particulate matter in internal combustion engines.71 The Ag@
CeO2 rice-ball system discussed in Section 7.2.1 was employed for
this purpose. Conventional Ag/CeO2 systems generally show low
activity below 300 °C and exhibit only limited thermal stability.28,72
The results reported for Ag@CeO2 indicated that the core-shell cata-
lyst showed good activity for soot oxidation at temperatures below

b1469_Ch-07.indd 385 4/8/2013 12:33:30 PM


b1469 Catalysis by Ceria and Related Materials

386 M. Cargnello, R. J. Gorte and P. Fornasiero

Figure 7.10 (a) Evaluation of catalytic activity for soot oxidation of rice-ball Ag@
CeO2 structures (squares), conventional Ag/CeO2 (diamonds), CeO2 alone (dots)
and in the absence of any catalyst (triangles). (b) Arrhenius plots. Reprinted from
Kayama et al.28 Copyright 2010 the American Chemical Society.

300 °C, while a conventional Ag/CeO2 system was active only for
temperatures well above this (Fig. 7.10).
Unfortunately, the thermal stability of the Ag@CeO2 was limited
because the ceria shell sintered at high temperatures (800 °C).
Indeed, the stability of similar structures prepared using a La-doped
ceria shell (La doping improves ceria thermal stability) was much
improved, to a sufficient extent that the catalyst prepared with
La doping exhibited activity similar to that shown by the rice-ball
catalyst in Fig. 7.11 before and after an aging treatment at 800 °C for
50 h.
Based on the Arrhenius plots in Fig. 7.10(b), the frequency fac-
tors of the rate constants for the reactions carried out over Ag@
CeO2, over CeO2, and with no catalyst were very similar. The primary
effect of the catalysts was to change the activation energy. This obser-
vation was explained by arguing that the formation of active oxygen
species was more efficient in the embedded system. Based on studies
of O2 isotopic exchange (IE), the authors argued that the highest
concentration of reactive oxygen species, which they believed to be
O2− based on electron spin resonance (ESR) measurements at 20 K,
were formed in the Ag@CeO2 sample.72 They argued that the

b1469_Ch-07.indd 386 4/8/2013 12:33:30 PM


b1469 Catalysis by Ceria and Related Materials

Core-Shell-Type Materials Based on Ceria 387

Figure 7.11 Mechanism for soot oxidation over a Ag@CeO2 catalyst with a rice-ball
morphology. Reprinted from Yamazaki et al.72 with permission from Elsevier.

improved activity of the rice-ball Ag@CeO2 morphology was due to


the presence of an appropriate Ag particle size and a large Ag–CeO2
interface for the production of these reactive oxygen species. They
believed the core-shell design is fundamentally important for achiev-
ing improved catalytic properties. They further suggested a reaction
mechanism in which active oxygen species are first formed at the
Ag–CeO2 interface (O∗) and then migrate through the CeO2 layer
(Onx−) to react with soot particles, as shown in Fig. 7.11.

7.3.3 Selective organic transformations


Ceria-supported metals are also good candidates for catalyzing
organic transformations. The main features that make ceria suitable
as a support for these reactions are its redox properties and the
coexistence of strong Lewis-base and weak Lewis-acid sites.73
The potentially advantageous, core-shell configuration of Ag–CeO2
composites was employed by Mitsudome et al. for chemoselective
reduction reactions.36 The main concept behind their work was that

b1469_Ch-07.indd 387 4/8/2013 12:33:30 PM


b1469 Catalysis by Ceria and Related Materials

388 M. Cargnello, R. J. Gorte and P. Fornasiero

they wanted to avoid the presence of Ag surfaces, which they


believed would catalyze the homolytic cleavage of H2 to H, which
according to their model would be indiscriminate as reducing
agents. Rather, they planned to take advantage of the Ag–CeO2
interface to produce heterolytically cleaved, active H species, which
they believed would be selective for the reduction of polar moieties
such as C = O. The catalytic activity of 10-nm Ag nanoparticles
embedded in 3–5 nm CeO2 particles was evaluated in the selective
reduction of nitrostyrenes to aminostyrenes and epoxide to alkenes
using molecular hydrogen. The authors were able to achieve good
selectivity. For example, in the case of 3-nitrostyrene, essentially
100% selectivity to 3-aminostyrene was achieved, with 98% yield.
The authors observed that the embedded Ag@CeO2 catalyst left the
C = C double bond completely intact. These results are impressive,
especially when the core-shell catalyst is compared to a conventional
Ag/CeO2 catalyst for which Ag nanoparticles of similar size were
deposited on a preformed CeO2 support. The conventional catalyst
showed a moderate yield for 3-nitrostyrene of 75%, and the hydro-
genation of the C = C double bond and other side-reactions were
competitive with respect to the reduction of the nitro group.
The Ag@CeO2 system is also active for the selective reduction of
other nitro compounds containing C = C double bonds, with the
reported yields and selectivity again being excellent. The same
authors demonstrated that these catalysts exhibit outstanding activ-
ity and selectivity for the deoxygenation of epoxides to the corre-
sponding alkenes and that they were recyclable by simple filtration.
There was no evidence for Ag leaching in the filtrates. Overall, the
results indicate that the core-shell structure resulted in novel cata-
lysts with different reaction properties and better stability against
metal leaching compared to conventional catalysts.
Xu and co-workers used plum-pudding multicore Pd@CeO2 and
Pt@CeO2 systems for selective oxidation reactions.31,32 The authors
demonstrated the photocatalytic oxidation of aromatic alcohols to
their corresponding aldehydes using oxygen at atmospheric pres-
sure and visible light. The yields observed in this study were not high
but the selectivities for the various substituted benzyl alcohols was

b1469_Ch-07.indd 388 4/8/2013 12:33:30 PM


b1469 Catalysis by Ceria and Related Materials

Core-Shell-Type Materials Based on Ceria 389

nearly 100%. This result is in sharp contrast with data for conven-
tional Pd/CeO2 catalysts, which displayed much lower conversions
and selectivities. The improved activity in the core-shell system was
again attributed to the formation of reactive oxygen species such as
superoxides.

7.3.4 Fuel-cell catalysts


The advantage of the core-shell configuration in limiting the sinter-
ing of the metal phase was demonstrated by our groups in fuel-cell
applications.74 Solid oxide fuel cells (SOFCs) are electrochemical
devices that can convert between electrical and chemical energy;
unlike low-temperature fuel cells, they are also capable of directly
using hydrocarbons as fuel.75 They typically operate at high tempera-
tures, generally around 700 °C, so that the use of thermally stable
materials is imperative. Although the fuel electrodes (the anodes) in
traditional SOFCs are based on bulk Ni composites with the electro-
lyte, there are significant advantages to using ceramic conductors in
place of Ni.76 With ceramic electrodes, it is necessary to add a cata-
lytic component in dopant amounts.
We took advantage of the dispersibility of Pd@CeO2 core-shell
structures39 to deposit them into the porous scaffold of SOFC mate-
rials as anode catalysts in order to enhance the thermal stability of
these materials. The porous scaffolds were composed of yttrium-
stabilized zirconia (YSZ) covered with a film of the conductive oxide
lanthanum strontium chromium manganite (LSCM). For compari-
son of the activity and thermal stability, we prepared other elec-
trodes that were identical except that the catalyst was simply Pd
(from Pd(II) nitrate) in one case and a mixture of Pd and CeO2
(from Pd(II) and Ce(III) nitrate salts) in the other. All the samples
were first calcined at 700 °C to remove any by-products and to stabi-
lize the materials. Then, accelerated aging tests were performed by
calcining the samples at 900 °C for 2 hours. Initially we tested all the
formulations in symmetric cells, e.g. cells where the anode and cath-
ode materials are the same. The corresponding Nyquist plots are
shown in Fig. 7.12(a).

b1469_Ch-07.indd 389 4/8/2013 12:33:30 PM


b1469 Catalysis by Ceria and Related Materials

390 M. Cargnello, R. J. Gorte and P. Fornasiero

Figure 7.12 (a) Nyquist plots of ac-impedance data obtained from symmetric cells
with LSCM/YSZ composite anodes containing the following catalysts: 1 wt% Pd; 1
wt% Pd + 9 wt% CeO2; 10 wt% Pd@CeO2. The circles correspond to data from
freshly prepared electrodes calcined at 700 °C (973 K) and the triangles correspond
to electrodes subjected to rapid thermal aging by calcining at 900 °C (1173 K).
(b) Maximum power density at 700 °C in humidified H2 (3% H2O) as a function of
time for complete fuel cells with Pd@CeO2 and Pd+CeO2 catalysts. Reproduced
from Kim et al.74 by permission of the Electrochemical Society.

The results showed that, in the case of samples calcined to 700 °C


(circles), an improvement was obtained by impregnation of both Pd
and CeO2 with respect to the Pd-only sample. This was in accord-
ance with the known properties of ceria in enhancing reaction rates
under redox conditions. The Pd@CeO2 formulation, however,
showed the best performance, with a value for the area specific
resistance (ASR) of only 0.11 Ω.cm2. The superiority of the embed-
ded system was further corroborated by the data after the aging
treatment. The Pd-only system showed a dramatic deactivation after
sintering at 900 °C, with an increase in its ASR of about 150%. The
Pd+CeO2 formulation was somewhat more stable, but still it dis-
played an increase of ASR of 60%. The Pd@CeO2 sample, on the
other hand, showed a very modest deactivation, with an increase in
ASR of only 18%. It was remarkable that the aged Pd@CeO2 elec-
trode was still more active than the two other formulations in their
“fresh” state.

b1469_Ch-07.indd 390 4/8/2013 12:33:30 PM


b1469 Catalysis by Ceria and Related Materials

Core-Shell-Type Materials Based on Ceria 391

Figure 7.13 SEM images of Pd+CeO2 catalyst supported on porous YSZ and
LSCM substrates after calcining at (a) 700 °C and (b) 900 °C. Bottom: Similar
investigation for a Pd@CeO2 formulation after calcining at (c) 700 °C and (d) 900 °C.
Reproduced from Kim et al.74 by permission of the Electrochemical Society.

To gain insights into the deactivation mechanism, we analyzed


the spent anodes using scanning electron microscopy (SEM) to
understand their morphology before and after use (Fig. 7.13).
SEM images of the Pd+CeO2 material before and after the aging
treatment are very different (Fig. 7.13(a) and (b)). Initially, the
fresh sample contained a film of CeO2 on top of the electrode mate-
rials with a few particles around 200 nm that were assigned to Pd
associated with CeO2 by EDS analysis. However, the number of par-
ticles was too low to account for the entire Pd and CeO2 loading,
and therefore a vast number of small Pd particles on top of the ceria
film must have been present but beyond instrument resolution.
Nevertheless, after the aging treatment several large particles were
clearly observed. Since the ceria film was stable under the present

b1469_Ch-07.indd 391 4/8/2013 12:33:31 PM


b1469 Catalysis by Ceria and Related Materials

392 M. Cargnello, R. J. Gorte and P. Fornasiero

conditions, these large particles were identified as Pd crystallites that


grew during the aging treatment by Ostwald ripening. SEM images
of the Pd@CeO2 formulation (Fig. 7.13(c) and (d)) both before and
after use showed the absence of any large agglomerate of particles,
thus demonstrating that the thermal stability of this sample was due
to suppressed sintering of the Pd phase because of the ceria shell
barrier.

7.4 Conclusions
This chapter gives an overview of core-shell materials based on
ceria. Some very recent developments have been presented to
highlight that these materials show novel properties compared
to conventional, supported ceria-based catalysts. The protection
offered by the ceria shell against sintering and the active role of the
metal/ceria interface under catalytic conditions are among the
most interesting properties of these systems. These characteristics
are a result of the special metal/support interaction that is cre-
ated as this interface is also maximized using the core-shell
configuration.
Though a very new field of research, this area is expected to
drastically increase in the next few years.

Acknowledgements
M.C. and P.F. acknowledge the University of Trieste and the Italian
Consortium on Materials Science and Technology (INSTM) for
financial support. R.J.G. acknowledges support from the Department
of Energy, Office of Basic Energy Sciences, Chemical Sciences,
Geosciences and Biosciences Division, Grant DE-FG02-85ER13350.

References
1. Schartl, W., Nanoscale, 2 (2010) 829–843.
2. Shemesh, Y., Macdonald, J.E., Menagen, G., Banin, U., Angew. Chem.,
Int. Ed., 50 (2011) 1185–1189.

b1469_Ch-07.indd 392 4/8/2013 12:33:31 PM


b1469 Catalysis by Ceria and Related Materials

Core-Shell-Type Materials Based on Ceria 393

3. Dong, A., Chen J., Vora, P.M., Kikkawa, J.M., Murray, C.B., Nature,
466 (2010) 474–477.
4. Greeley, J., Norskov, J.K., Kibler, L.A., El-Aziz, A.M., Kolb, D.M.,
J. Chem. Phys., 7 (2006) 1032–1035.
5. Tao, F., Grass, M.E., Zhang, Y., Butcher, D.R., Renzas, J.R., Liu, Z.,
Chung, J.Y., Mun, B.S., Salmeron, M., Somorjai, G.A., Science, 322
(2008) 932–934.
6. Lim, D.K., Jeon, K.S., Hwang, J.H., Kim, H., Kwon, S., Suh, Y.D., Nam,
J.M., Nat. Nanotechnol., 6 (2011) 452–460.
7. Yeung, C.M.Y., Meunier, F., Burch, R., Thompsett, D., Tsang, S.C.,
J. Phys. Chem. B, 110 (2006) 8540–8543.
8. Fu, Q., Saltsburg, H., Flytzani-Stephanopoulos, M., Science, 301 (2003)
935–938.
9. Wang, X., Gorte, R.J., Wagner, J.P., J. Catal., 212 (2002) 225–230.
10. De Rogatis, L., Cargnello, M., Gombac, V., Lorenzut, B., Montini, T.,
Fornasiero, P., ChemSusChem, 3 (2010) 24–42.
11. Bunluesin, T., Gorte, R.J., Graham, G.W., Appl. Catal. B, 15 (1998)
107–114.
12. Roh, H.S., Jun, K.W., Dong, W.S., Chang, J.S., Park, S.E., Joe, Y.I.,
J. Mol. Catal. A: Chem., 181 (2002) 137–142.
13. Wang, X. Gorte, R.J., Appl. Catal. A, 224 (2002) 209–218.
14. Montoya, J.A., Romero-Pascual, E., Gimon, C., Del Angel, P.,
Monzón, A., Catal. Today, 63 (2000) 71–85.
15. Wang, S., Lu, G.Q., Appl. Catal. B, 19 (1998) 267–277.
16. Trovarelli, A., Catal. Rev. Sci. Eng., 38 (1996) 439–520.
17. Bernal, S., Calvino, J.J., Cauqui, M.A., Gatica, J.M., Larese, C.,
Pérez Omil, J.A., Pintado, J.M., Catal. Today, 50 (1999) 175–206.
18. Gorte, R.J., Vohs, J.M., Annu. Rev. Chem. Biomol. Eng., 2 (2011) 9–30.
19. Green, I.X., Tang, W., Neurock, M., Yates, J.T., Angew. Chem., Int. Ed.,
50 (2011) 10186–10189.
20. Green, I.X., Tang, W., Neurock, M., Yates, J.T., Science, 333 (2011)
736–739.
21. Arnal, P.M., Comotti, M., Schuth, F., Angew. Chem., Int. Ed., 45 (2006)
8224–8227.
22. Park, J.N., Forman, A.J., Tang, W., Cheng, J., Hu, Y.S., Lin, H.,
McFarland, E.W., Small, 4 (2008) 1694–1697.

b1469_Ch-07.indd 393 4/8/2013 12:33:31 PM


b1469 Catalysis by Ceria and Related Materials

394 M. Cargnello, R. J. Gorte and P. Fornasiero

23. Joo, S.H., Park, J.Y., Tsung, C.K., Yamada, Y., Yang, P., Somorjai, G.A.,
Nat. Mater., 8 (2009) 126–131.
24. Lee, I., Joo, J.B., Yin, Y., Zaera, F., Angew. Chem., Int. Ed., 50 (2011)
10208–102011.
25. Cushing, B.L., Kolesnichenko, V.L., O’Connor, C.J., Chem. Rev., 104
(2004), 3893–3946.
26. Huang, Y., Wang, A., Li, L., Wang, X., Su, D., Zhang, T., J. Catal., 255
(2008) 144–152.
27. Cargnello, M., Gentilini, C., Montini, T., Fonda, E., Mehraeen, S.,
Chi, M., Herrera-Collado, M., Browning, N.D., Polizzi, S., Pasquato, L.,
Fornasiero, P., Chem.Mater., 22 (2010) 4335–4345.
28. Kayama, T., Yamazaki, K., Shinjoh, H., J. Am. Chem. Soc., 132 (2010)
13154–13155.
29. Cargnello, M., Montini, T., Polizzi, S., Wieder, N.L., Gorte, R.J.,
Graziani, M., Fornasiero, P., Dalton Trans., 39 (2010) 2122–2127.
30. Hwai, P.C., Kong, Y.L., Liu, H., J. Mater. Chem., 12 (2002) 934–937.
31. Zhang, N., Fu, X., Xu, Y.J., J. Mater. Chem., 21 (2011) 8152–8158.
32. Zhang, N., Liu, S., Fu, X., Xu, Y.J., J. Phys. Chem. C, 115 (2011)
22901–22909.
33. Ganguli, A.K., Ganguly, A., Vaidya, S., Chem. Soc. Rev., 39 (2010)
474–485.
34. Yeung, C.M.Y., Yu, K.M.K., Fu, Q.J., Thompsett, D., Petch, M.I.,
Tsang, S.C., J. Am. Chem. Soc., 127 (2005) 18010–18011.
35. Yeung, C.M.Y., Tsang, S.C., J. Phys. Chem. C, 113 (2009) 6074–6087.
36. Mitsudome, T., Mikami, Y., Matoba, M., Mizugaki, T., Jitsukawa, K.,
Kaneda, K., Angew. Chem., Int. Ed., 51 (2012) 136–139.
37. Liz-Marzán, L.M., Giersig, M., Mulvaney, P., Langmuir, 12 (1996)
4329–4335.
38. Jana, N.R., Earhart, C., Ying, J.Y., Chem. Mater., 19 (2007) 5074–5082.
39. Cargnello, M., Wieder, N.L., Montini, T., Gorte, R.J., Fornasiero, P.,
J. Am. Chem. Soc., 132 (2010) 1402–1409.
40. Cargnello, M., Wieder, N.L., Canton, P., Montini, T., Giambastiani, G.,
Benedetti, A., Gorte, R.J., Fornasiero, P., Chem. Mater., 23 (2011)
3961–3969.
41. Gradeff, P.S., Schreiber, F.G., Brooks, K.C., Sievers, R.E., Inorg. Chem.,
24 (1985) 1110–1111.

b1469_Ch-07.indd 394 4/8/2013 12:33:31 PM


b1469 Catalysis by Ceria and Related Materials

Core-Shell-Type Materials Based on Ceria 395

42. Trimm, D.L., Appl. Catal. A, 296 (2005) 1–11.


43. Rozovskii, A.Y. Lin, G.I., Top. Catal., 22 (2003) 137–150.
44. Davda, R.R., Dumesic, J.A., Angew. Chem., Int. Ed., 42 (2003) 4068–4071.
45. Velu, S., Suzuki, K., Kapoor, M.P., Ohashi, F., Osaki, T., Appl. Catal. A,
213 (2001) 47–63.
46. De Rogatis, L., Montini, T., Casula, M.F., Fornasiero, P., J. Alloys
Compd., 451 (2008) 516–520.
47. Cortright, R.D., Davda, R.R., Dumesic, J.A., Nature, 418 (2002)
964–967.
48. Twigg, M.V., Spencer, M.S., Appl. Catal. A, 212 (2001) 161–174.
49. Fu, Q., Deng, W., Saltsburg, H., Flytzani-Stephanopoulos, M., Appl.
Catal. B, 56 (2005) 57–68.
50. Li, Y., Fu, Q., Flytzani-Stephanopoulos, M., Appl. Catal. B, 27 (2000)
179–191.
51. Hurtado-Juan, M.A., Yeung, C.M.Y., Tsang, S.C., Catal. Commun.,
9 (2008) 1551–1557.
52. Yeung, C.M.Y., Tsang, S.C., J. Mol. Catal. A: Chem., 322 (2010) 17–25.
53. Yeung, C., Tsang, S., Catal. Lett., 128 (2009) 349–355.
54. Shido, T., Iwasawa, Y., J. Catal., 141 (1993) 71–81.
55. Shido, T., Asakura, K., Iwasawa, Y., J. Catal., 122 (1990) 55–67.
56. Ruettinger, W., Liu, X., Farrauto, R.J., Appl. Catal. B, 65 (2006)
135–141.
57. Lieske, H., lter, J., J. Phys. Chem., 89 (1985) 1841–1842.
58. Nagai, Y., Dohmae, K., Ikeda, Y., Takagi, N., Tanabe, T., Hara, N.,
Guilera, G., Pascarelli, S., Newton, M.A., Kuno, O., Jiang, H., Shinjoh, H.,
Matsumoto, S., Angew. Chem., Int. Ed., 47 (2008) 9303–9306.
59. Newton, M.A., Belver-Coldeira, C., Martínez-Arias, A., Fernández-
García, M., Angew. Chem., Int. Ed., 46 (2007) 8629–8631.
60. Wieder, N.L., Cargnello, M., Bakhmutsky, K., Montini, T., Fornasiero, P.,
Gorte, R.J., J. Phys. Chem. C, 115 (2010) 915–919.
61. Da Silva, J.L.F., Ganduglia-Pirovano, M.V., Sauer, J., Bayer, V.,
Kresse, G., Phys. Rev. B: Condens. Matter, 75 (2007) 089901.
62. Da Silva, J.L.F., Ganduglia-Pirovano, M.V., Sauer, J., Phys. Rev.
B: Condens. Matter, 76 (2007) 045121.
63. Zhou, G., Shah, P.R., Montini, T., Fornasiero, P., Gorte, R.J., Surf. Sci.,
601 (2007) 2512–2519.

b1469_Ch-07.indd 395 4/8/2013 12:33:31 PM


b1469 Catalysis by Ceria and Related Materials

396 M. Cargnello, R. J. Gorte and P. Fornasiero

64. Gorte, R.J., AIChE J., 56 (2010) 1126–1135.


65. Roberts, S., Gorte, R.J., J. Catal., 124 (1990) 553–556.
66. Mehta, V., Cooper, J.S., J. Power Sources, 114 (2003) 32–53.
67. Bion, N., Epron, F., Moreno, M., Marino, F., Duprez, D., Top. Catal.,
51 (2008) 76–88.
68. Lin, J., Huang, Y., Li, L., Qiao, B., Wang, X., Wang, A., Zhang, T.,
Chem. Eng. J., 168 (2011) 822–826.
69. Lin, J., Li, L., Huang, Y., Zhang, W., Wang, X., Wang, A., Zhang, T.,
J. Phys. Chem. C, 115 (2011) 16509–16517.
70. Zanella, R., Giorgio, S., Henry, C.R., Louis, C., J. Phys. Chem. B, 106
(2002) 7634–7642.
71. Stanmore, B.R., Brilhac, J.F., Gilot, P., Carbon, 39 (2001) 2247–2268.
72. Yamazaki, K., Kayama, T., Dong, F., Shinjoh, H., J. Catal., 282 (2011)
289–298.
73. Vivier, L., Duprez, D., ChemSusChem, 3 (2010) 654–678.
74. Kim, J.S., Wieder, N.L., Abraham, A.J., Cargnello, M., Fornasiero, P.,
Gorte, R.J., Vohs, J.M., J. Electrochem. Soc., 158 (2011) B596–600.
75. Park, S., Vohs, J.M., Gorte, R.J., Nature, 404 (2000) 265–267.
76. Gorte, R.J., Vohs, J.M., Curr. Opin. Colloid Interface Sci., 14 (2009)
236–244.
77. Bakhmutsky, K., Zhou, G., Timothy, S., Gorte, R. J. Catal. Lett. 129
(2009) 61–65.

b1469_Ch-07.indd 396 4/8/2013 12:33:31 PM


b1469 Catalysis by Ceria and Related Materials

CHAPTER 8

NEW DEVELOPMENTS IN CERIA-BASED


MIXED OXIDE SYNTHESIS
AND REACTIVITY IN COMBUSTION
AND OXIDATION REACTIONS
Benjaram M. Reddy, Thallada Vinod Kumar
and Naga Durgasri
Inorganic and Physical Chemistry Division,
Indian Institute of Chemical Technology, Hyderabad, 500 607, India

8.1 Introduction
There has been an explosion of interest and investment in nanosci-
ence and nanotechnology over the last few years. The nanoscience
revolution is one of the biggest things to happen since the beginning
of modern science, and it is nowadays at the centre of future techno-
logical progress owing to the increasing ability to manipulate matter
on the nanometre scale. One of the important driving forces for the
rapidly developing field of nanoparticle synthesis is the distinctly
different physicochemical properties exhibited by nanoparticles
compared to their bulk counterparts.1–3 The differences may be due
to surface effects, small size effects, quantum size effects and so on;
they open up new opportunities for the development of materials
with unusual or tailored properties. Like nanomaterials, bulk materi-
als also exhibit surface-dependent properties but these are dominant
in the case of nanoparticles only because they possess a vast surface
area per unit volume and a high proportion of atoms at the surface
and near surface layers rather than in the particle interior.

397

b1469_Ch-08.indd 397 4/8/2013 12:34:53 PM


b1469 Catalysis by Ceria and Related Materials

398 B. M. Reddy, T. Vinod Kumar and N. Durgasri

Many properties of nanoparticles are directly connected to their


small size. The small size leads to many distinct properties, which
influences lattice symmetry and cell parameters. They display inter-
esting optical properties since the optical properties are dependent
on the absorption and emission wavelengths, which can be con-
trolled by particle size and surface functionalization. The chemical
nature and the size of the nanoparticles control the ionic potential
or the electron affinity and thereby the electron transport proper-
ties. Nanomaterials have numerous applications in catalysis, elec-
tronics, photovoltaics, etc. The application prospects for catalysts
are enhanced by their high surface area per unit volume and the
homogeneous distribution of the nanoparticles.4,5
For the control of environmental pollution, catalysts or sorbents
that contain redox metal oxides are normally employed. Among
them, ceria exhibits unique features due to its defect crystal struc-
ture involving facile transformations between the Ce3+ and Ce4+
oxidation states. This unique redox process can be triggered either
by direct electron transfer from a metal particle (or molecule)
adsorbed over the ceria surface, which would lead to partially posi-
tively charged (cationic) adsorbates, or by oxygen release leaving an
oxygen vacancy and two electrons localized on a Ce cation, which
becomes Ce3+. The former effect is particularly important when
dealing with metals supported on ceria. The latter process is related
to the ability of ceria to act as an oxygen buffer (referred to as the
oxygen storage/release component). A crucial role is assigned to
surface oxygen vacancies, which are likely to form more readily at
the nanoscale. Due to this property ceria is a key component mate-
rial in auto-exhaust purification three-way catalytic converters
(TWCs) and applications such as fuel cells, optical materials, gas
sensors and so forth.6–9 Recently, ceria has also been the focus of
research on the health effects of nanoparticles.10 The application of
ceria on its own is limited to low-temperature regimes. For high-
temperature applications, the loss of the oxygen storage/release
capacity (OSC) is a major disadvantage. The loss is due to sintering,
which does not allow ceria to switch rapidly between 3+ and 4+ oxi-
dation states. In addition, the poor low-temperature performance of

b1469_Ch-08.indd 398 4/8/2013 12:34:54 PM


b1469 Catalysis by Ceria and Related Materials

New Developments in Ceria-Based Mixed Oxide Synthesis 399

the material means that ceria-only materials cannot meet the


increasingly demanding standards required by legislation for auto-
motive pollution-control devices. This has forced researchers to seek
new catalyst compositions with the aim of: (i) increasing thermal
stability and (ii) enhancing low-temperature redox performance.
Therefore, a great deal of attention has been paid to doping ceria
with isovalent or aliovalent cations to improve OSC and thermal
stability.
Properties of materials are largely dependent on size and mor-
phology, so control over synthetic methodologies is important. This
is because the growth of the materials in nanoscale is largely
dependent on the thermodynamic and kinetic barriers in reactions,
as defined by the reaction trajectory and influenced by vacancies,
defects and surface reconstruction.11 The simultaneous control of
particle size and shape together with their uniformity is one of the
key objectives in many synthetic procedures. In the present chapter,
new developments in various preparation methodologies for the
fabrication of ceria-based materials are addressed. The reactivity of
ceria-based nano-oxides for non-selective oxidation or catalytic com-
bustion or oxidation reactions is also discussed.

8.2 Synthetic Approaches


Synthetic methods of preparing materials are broadly divided into
two categories: physical and chemical methods, and include co-
precipitation, hydrothermal and solvothermal, sol-gel, microemul-
sion, microwave, sonochemical, impregnation, combustion, ball
milling and so on. There are several parameters, which influence
the size and morphology of the finished material.12 By controlling
these parameters it is possible to make well-defined ceria-based
oxides as shown in Fig. 8.1.
The chemical reactions that take place in the liquid phase and
the formation of high-quality nanoparticles (that is, those with a
defined diameter, almost monodisperse size distribution and small
degree of agglomeration) are in general limited by two major
boundary conditions: firstly, the precise and directed control of

b1469_Ch-08.indd 399 4/8/2013 12:34:54 PM


b1469 Catalysis by Ceria and Related Materials

400 B. M. Reddy, T. Vinod Kumar and N. Durgasri

Figure 8.1 Synthesis of ceria-based oxides. Reprinted with permission from Guo
and Wang.12 Copyright 2011 the American Chemical Society.

nucleation and nucleus growth and, secondly, the effective suppres-


sion of agglomeration processes. The growth and stabilization of
nanoparticles is schematically shown in Fig. 8.2.
The growth or agglomeration of nanoparticles can be decreased
by electrostatic stabilization or steric stabilization. Electrostatic stabi-
lization can be achieved by targeted adsorption of ions (H+, OH−,
SO42−, etc.) on the particle surface. Steric stabilization is achieved by
adsorption of long-chain organic molecules (e.g., oleyl amine, oleic
acid, trioctyl phosphine, trioctyl phosphine oxide, etc.).4

8.2.1 Co-precipitation
In recent decades, co-precipitation was recognized as a method of
preparing nanomaterials using a trial-and-error approach. Often

b1469_Ch-08.indd 400 4/8/2013 12:34:54 PM


b1469 Catalysis by Ceria and Related Materials

New Developments in Ceria-Based Mixed Oxide Synthesis 401

Figure 8.2 Growth and stabilization of nanoparticles. Reprinted with permission


from Helmut and Claus.4 Copyright 2010 Wiley-VCH.

in this method, the salt compounds of two desired precursors are


dissolved in aqueous solvents and subsequently precipitated with a
precipitant by pH adjustment depending on the material. There
are three mechanisms involved in co-precipitation: inclusion,
occlusion and adsorption. In any method of inducing precipita-
tion, supersaturation is necessary. Principally, particle size and the
morphology of the products is determined by nucleation and
growth. These are mainly influenced by the concentration of the
precursors, temperature, rate of addition of precipitating agent
(urea, hexamethylenetetramine, NaOH, NH4OH, Na2CO3, etc.)
and the pH of the medium. To achieve high homogeneity, the
solubility products of the precipitates of the metal cations must be
close because the difference in solubility product affects the pre-
cipitation kinetics. Co-precipitation results in atomic scale mixing
and hence the calcination temperature required for the formation
of the final product is low, which produces small particles.13–16
Typical advantages associated with this method are (i) artless and
hasty preparation, (ii) the particle size can be adjustable and (iii)

b1469_Ch-08.indd 401 4/8/2013 12:34:54 PM


b1469 Catalysis by Ceria and Related Materials

402 B. M. Reddy, T. Vinod Kumar and N. Durgasri

(a) (b) (c)

Figure 8.3 (a) TEM and (b), (c) HREM images of a ceria–lanthana sample
calcined at 1,073 K. Inset: Enlarged views of selected areas. Reprinted with permis-
sion from Reddy et al.19 Copyright 2010 the American Chemical Society.

flexibility in modifying the particle surface state and overall


homogeneity.15
By exploiting these advantages, a series of Ce1−xMxO2−δ (M = Zr4+,
Hf , La3+, Tb3+/4+, Pr3+/4+, etc.) nano-oxides were synthesized by
4+

Reddy and co-investigators.17–21 Transmission electron microscopy


(TEM) and high resolution electron microscopy (HREM) images of
ceria–lanthana (CL) are shown in Fig. 8.3. The materials prepared
were thoroughly characterized by different techniques and investi-
gated for various catalytic reactions, and proven to be superior in
terms of both thermal and textural behaviour as well as catalytic activ-
ity (OSC and CO oxidation). Among them, the Ce0.8Hf0.2O2 combina-
tion exhibited a high OSC and good CO oxidation activity.22
Using conventional co-precipitation control of particle size is
often difficult. To overcome the problem, various modifications
have been introduced, which include redox co-precipitation,
carbonate co-precipitation, co-precipitation with supercritical dry-
ing, template-assisted co-precipitation, inverse co-precipitation, etc.
Scanning electron microscopy (SEM) images of a non-porous MgO–
CeO2 material obtained by one of these approaches are shown in
Fig. 8.4. To obtain this material, Abimanyu et al.23 used template-
assisted co-precipitation. In this approach, an ionic liquid (IL) was
used as a template. The IL has a low vapour pressure, which helps
in reducing shrinkage during aging and drying and prevents reduc-
tion of the surface area. The IL contains an anion part and a cation
part, which are important in determining the pore size and pore

b1469_Ch-08.indd 402 4/8/2013 12:34:54 PM


b1469 Catalysis by Ceria and Related Materials

New Developments in Ceria-Based Mixed Oxide Synthesis 403

Figure 8.4 SEM images of MgO–CeO2 catalysts prepared by co-precipitation (a)


with IL BBF4 and (b) without IL. Reprinted with permission from Haznan et al.23
Copyright 2007 the American Chemical Society.

volume of the materials. It was observed that when the catalysts were
prepared using ILs, the surface area and pore volume were
enhanced 2.0–6.0 and 5.5–10.7 times, respectively, compared to
materials prepared by simple co-precipitation without ILs. The use
of ILs in the co-precipitation method changes not only the structure
of the particles, but also surface basicity and basic strength
distribution.
To prepare ceria–zirconia over alumina, Wang et al.24 used two
methods: co-precipitation with supercritical drying and co-precipita-
tion with common drying. When drying was performed under
supercritical conditions in ethanol (250 °C, 7.5 MPa) there was no
liquid/vapour interface during drying, which resulted in smaller
crystallites, high BET surface area, better thermal stability and supe-
rior redox properties compared to the material prepared by co-
precipitation with common drying. In another study, Feng et al.25
prepared cubic fluorite nano-Ce1-xZrxO2 with co-precipitation using
NH4OH as precipitant followed by hydrothermal crystallization at
different temperatures.
By adopting a similar approach, Reddy et al.26 prepared alumina-
supported nanosized ceria–terbia solid solutions by deposition co-
precipitation at ambient conditions. The literature reveals several

b1469_Ch-08.indd 403 4/8/2013 12:34:58 PM


b1469 Catalysis by Ceria and Related Materials

404 B. M. Reddy, T. Vinod Kumar and N. Durgasri

investigations on the preparation of ceria-based materials by the


simplest co-precipitation method. Typical disadvantages associated
with this method are that: (i) it is not suitable for the preparation of
high purity and accurate stoichiometric phases, (ii) it does not work
well if the reactants have very different solubility as well as different
precipitation rates and (iii) there are no universal experimental
conditions for the synthesis of various types of metal oxides.

8.2.2 Hydrothermal method


Among the various branches of science, the word hydrothermal is
from the earth sciences.27,28 It refers to the synthesis of oxide materi-
als by chemical reactions using water as a reaction medium in a
sealed heated vessel generally in an autoclave or a bomb calorime-
ter-type apparatus above 100 °C. The hydrothermal method is one
of the most important and well-established methods for the labora-
tory and industrial scale synthesis of nanomaterials. The production
of nanomaterials using an autoclave has many advantages. Most
importantly, the operational temperatures are well below the melt-
ing point of the reactants and the reactivity of inorganic solids is
significantly affected due to the solubility of various compounds
under elevated temperatures and pressures, whereas the behaviour
of water (the solvent) is different. Different types of autoclaves and
tunable reaction parameters make this method very useful.29–31
Because reactions are performed in closed systems pollution is mini-
mized, reagents can be recycled and energy consumption is low.29
The mechanism is different from that of solid-state reactions. The
former follows the liquid nucleation model and in the latter the
mechanism is governed by the diffusion of atoms or ions at the
interface between reactants. Kinetically stable phases are favoured
in short-time processes while thermodynamically stable phases are
generally formed in long-term experiments — depending on the
chosen temperature-pressure system.32–34
Qin et al.35 prepared CuO–CeO2 nanospheres by a hydrothermal
method and evaluated them for CO oxidation. Previously, the lowest
CO oxidation temperature reported for CuO–CeO2 catalysts was

b1469_Ch-08.indd 404 4/8/2013 12:34:58 PM


b1469 Catalysis by Ceria and Related Materials

New Developments in Ceria-Based Mixed Oxide Synthesis 405

Figure 8.5 SEM images of as-prepared CeCu0.33 nanospheres at (a) low magnifica-
tion and (b) high magnification. Reprinted from Qin et al.35 with permission from
the Royal Society of Chemistry.

100 °C, and with this method it decreased to 71 °C. In this proce-
dure, Ce(NO3)3.6H2O and Cu(NO3)2.3H2O were used as the precur-
sors and urea as the precipitating agent. The prepared solutions
were transferred to a Teflon-lined autoclave and heated at 180 °C for
100 min, which led to the precipitation of the precursor. The pre-
cipitate was centrifuged, washed, dried and finally calcined at 600 °C
for 4 h. SEM images of the resulting products are shown in Fig. 8.5.
The control of shape and particle size is not easy with this
method. Therefore, various modifications have been introduced to
improve the material properties and include the surfactant-assisted
hydrothermal method, the microwave-assisted hydrothermal
method, the supercritical hydrothermal method, etc.
There is a relation between material properties and morphology.
In order to investigate this relation, Zhang et al.36 prepared nanorod-
like, micro-spherical, micro-bowknot-like and micro-octahedral
shaped ceria–zirconia–yttria (CZY) solid solutions with the surfactant-
assisted hydrothermal method using a triblock copolymer (Pluronic
P123) or cetyltrimethylammonium bromide (CTAB) surfactant. The
formation mechanism of these materials by the surfactant-assisted
hydrothermal method is shown schematically in Fig. 8.6.
To prepare different shapes of CZY particles, firstly the precipita-
tion mixture was obtained by mixing metal precursors [Ce(NO3)3·
6H2O, ZrO(NO3)2·2H2O and Y(NO3)3·6H2O], a surfactant and urea.

b1469_Ch-08.indd 405 4/8/2013 12:34:58 PM


b1469 Catalysis by Ceria and Related Materials

406 B. M. Reddy, T. Vinod Kumar and N. Durgasri

Figure 8.6 Formation of ceria–zirconia-yttria precursors and solid solutions under


surfactant-assisted hydrothermal conditions. Reprinted with permission from
Zhang et al.36 Copyright 2009 the American Chemical Society.

The precipitation mixture was placed in an oven for aging at 80 °C


for 72 h. The milky slurry produced was transferred into a Teflon-
lined autoclave for hydrothermal treatment at different temperatures
(100, 120 or 240 °C for 48 or 72 h). After being cooled to room tem-
perature the solution was filtered, washed with deionized water, dried
overnight at 60 °C and calcined in an air flow. Samples derived from
the CTAB-assisted hydrothermal treatment at 120 °C for 72 h are
denoted as CZY-CTAB-120, whereas those derived from the P123-
assisted hydrothermal treatment at 100, 120 and 240 °C for 48 h are
denoted as CZY-P123–100, CZY-P123–120 and CZY-P123–240, respec-
tively. The reducibility of these nanomaterials and micromaterials at
low temperatures (240–550 °C) was enhanced in the order: micro-
octahedral CZY < micro-spherical CZY < micro-bowknot-like CZY <
nanorod-like CZY.36
The supercritical state is defined as the state at which the distinc-
tion between the liquid state and the gaseous state disappear and the
material exists as a fluid; it occurs above the critical temperature and
critical pressure. For water, the critical temperature and pressure are
373 °C and 22.1 MPa, respectively. In this state, various fluid proper-
ties change, including the density of the molecules, viscosity, diffu-
sion coefficient and thermal conductivity. The dielectric constant of

b1469_Ch-08.indd 406 4/8/2013 12:34:58 PM


b1469 Catalysis by Ceria and Related Materials

New Developments in Ceria-Based Mixed Oxide Synthesis 407

water decreases considerably with increasing temperature. In the


hydrothermal method, the dielectric constant controls the reaction
rate, equilibrium and solubility of metal oxides. The solubility of a
metal oxide first increases with an increase in temperature and then
decreases; this decrement is due to the decrease of density and die-
lectric constant. According to nucleation theory, the nucleation rate
is expressed as a function of the degree of supersaturation and the
surface energy. A decrease in solubility leads to a high degree of
supersaturation; an extremely high nucleation rate is expected
under supercritical conditions leading to the formation of small
nanoparticles of metal oxides.37 By using this idea, Kim et al.38 pre-
pared ceria–zirconia mixed oxides by continuous hydrothermal
synthesis in supercritical water. Several groups exploited this meth-
odology to prepare metal oxides for various applications.

8.2.3 Solvothermal method


The solvothermal method is similar to hydrothermal synthesis
except for the solvent used. In this procedure, non-aqueous solvents
(organic solvents) are used as the reaction medium. The use of
organic solvents has been reported to be an effective approach to
avoid the oxidation, hydrolysis and volatilization of nano-oxides and
their reactants.34 This method was advantageously exploited to make
nanosized ceria–zirconia mixed oxides.39
Fernández-González et al.40 synthesized nanostructured praseo-
dymium-doped ceria–titania mixed oxides by the solvothermal
method. They used Ce(NO3)6(NH4)2, Pr(NO3)3 and tetra isopropyl
orthotitanate as metal precursors, ethanol as the solvent and acety-
lacetone (acac) as the complexing agent. The complexing agent
controls the hydrolysis and condensation processes, and helps the
stabilization of the heterometallic nanocrystals in organic solvents.
The powder obtained was finally calcined. SEM images of the syn-
thesized samples are shown in Fig. 8.7.
The solvothermal reaction under supercritical conditions is
suitable for the synthesis of functional materials, since the reaction
rate and equilibrium can be varied by changing the dielectric

b1469_Ch-08.indd 407 4/8/2013 12:34:59 PM


b1469 Catalysis by Ceria and Related Materials

408 B. M. Reddy, T. Vinod Kumar and N. Durgasri

Figure 8.7 SEM images of the as-synthesized (raw) Ce0.45Pr0.05Ti0.5O2 sample and
after firing at 1000 °C. Reprinted from Fernández-González et al.40 with permission
from the Royal Society of Chemistry.

constant and the density of the solvents varies with temperature


and pressure. Devaraju et al.41 produced rod- and sphere-like CeO2
particles via a supercritical solvothermal method using CeCl3·7H2O
or Ce(NO3)3·6H2O as cerium precursors and ethanol or methanol
as solvents. The effect of reaction time on the morphology was
investigated. Reactions were performed at 400 °C for 15 min and
500 °C for 1 h followed by calcination in air. Field emission scan-
ning electron microscopy (FESEM) images of the resulting samples
are shown in Figs. 8.8 and 8.9, respectively. It was observed that
the samples prepared in different solvents exhibit different parti-
cle sizes due to different dielectric constants, which cause a

b1469_Ch-08.indd 408 4/8/2013 12:34:59 PM


b1469 Catalysis by Ceria and Related Materials

New Developments in Ceria-Based Mixed Oxide Synthesis 409

Figure 8.8 FESEM images of as-prepared samples using CeCl3·7H2O in (a) etha-
nol and (c) methanol at 400 °C for 15 min, and calcined in air at 500 °C for 1 h,
(b) and (d), respectively. Reprinted with permission from Devaraju et al.41 Copyright
2009 the American Chemical Society.

difference in the solubility of the starting materials under super-


critical conditions.

8.2.4 Sol-gel method


In recent decades, several studies have been made of sol-gel chemis-
try. This method has been widely used in the ceramics industry and
materials science. Typically metal alkoxides and metal chloride pre-
cursors are used as the starting materials. The reaction is very easy
to perform and does not require any special conditions. It is highly
suitable for the preparation of metal oxides. The name itself implies

b1469_Ch-08.indd 409 4/8/2013 12:34:59 PM


b1469 Catalysis by Ceria and Related Materials

410 B. M. Reddy, T. Vinod Kumar and N. Durgasri

Figure 8.9 FESEM images of as-prepared samples using Ce(NO3)3·6H2O in


(a) ethanol and (c) methanol at 400 °C for 15 min, and calcined in air at 500 °C
for 1 h, (b) and (d), respectively. Reprinted with permission from Devaraju et al.41
Copyright 2009 the American Chemical Society.

the development of inorganic networks through the formation of a


colloidal suspension (sol) and the gelation of the sol to form a net-
work in a continuous liquid phase (gel).42,43 The size of a sol particle
depends on the solution composition, temperature, pH, etc. The
physicochemical properties of the end product mainly depend on
the hydrolysis, condensation and drying conditions. The rate of
hydrolysis and condensation are greatly influenced by the type of
precursor, the ratio between alkoxide and water, the concentration
of acid and base, temperature, pH, type of solvent, electronegativity
of the metal atoms and so on.44,45 The quality of the gel46,47 depends
on the relative rates of hydrolysis and condensation processes, which
are given in Table 8.1. For different combinations of metal oxides,

b1469_Ch-08.indd 410 4/8/2013 12:34:59 PM


b1469 Catalysis by Ceria and Related Materials

New Developments in Ceria-Based Mixed Oxide Synthesis 411

Table 8.1 Effect of relative rates of hydrolysis and condensation on gel quality.
Reproduced with permission from David et al.47 Copyright 1995 the American
Chemical Society.
Hydrolysis rate Condensation rate Result
Slow Slow Colloids/sol
Slow Fast Controlled precipitation
Fast Fast Colloidal gel or gelatinous precipitate
Fast Slow Polymeric gels

the morphology of the resulting network depends on the chemical


reactivity of the metal oxides. A large difference in reactivity can
often cause phase separation.43,13
Drying the gel is another important step in the process, because
when liquid is removed from the gel several changes may occur. In
general, there are two types of drying processes: conventional drying
and supercritical drying. Normal drying of the gel leads to structural
collapse due to capillary forces drawing the walls of the pores
together thus reducing the pore size. The dried sample, known as a
xerogel, thus often has a surface area and pore volume that is too low
to be of catalytic interest. During supercritical drying, there is no
liquid/vapour interface and thus no capillary force; as a result, there
is no capillary-force-driven collapse or shrinkage of the highly open
gel networks. Such sol-gel materials are referred to as aerogels.13,45,48
A variety of materials can be synthesized by the sol-gel process, which
is schematically shown in Fig. 8.10.49 The main advantage of the sol-
gel approach is the flexibility of the sol-gel chemistry and the wide
range of microstructures that can be attained from nanoparticles,
nanostructured or nanoporous films to nanostructured monoliths.50
Auroux and co-investigators51 prepared Me2O3–CeO2 (Me = B,
Al, Ga, In) mixed oxides by the sol-gel method. The precursors for
group III metal oxides were trimethylborate, aluminium triisopro-
poxide and indium(III) acetylacetonate. The hydrolysis was per-
formed using a solution of ammonia and water. Of these materials,
a high specific surface area and good dispersion were achieved for
Al2O3–CeO2.

b1469_Ch-08.indd 411 4/8/2013 12:34:59 PM


b1469 Catalysis by Ceria and Related Materials

412 B. M. Reddy, T. Vinod Kumar and N. Durgasri

Figure 8.10 Synthesis of different materials using the sol-gel method.49

Tao et al.52 produced Ce0.9Gd0.1O1.95 (gadolinium-doped ceria,


GDC) nanopowders by a novel sol-gel thermolysis method using a
combination of citric acid and polyvinylpyrrolidone (PVP). In this
method, primarily the metal precursors, HNO3 and citric acid were
mixed. A citrate–metal ion complex was formed and different
amounts of surfactants were added to the complex to achieve a well-
dispersed, homogeneous GDC nanopowder. The PVP acts not only
as a fuel but also affects the gelation process and the aggregation of
particles during gel thermolysis due to its long chain structure.
Samples were derived using different surfactant contents and heat
treated at 600 °C for 2 h. SEM images of various GDC powders are
shown in Fig. 8.11.
The effect of different amounts of surfactant on the morphology
of the resultant powders was investigated. Different amounts of PVP
as a surfactant are denoted as S1 (2 × 10−3 g.ml−1), S2 (4 × 10−3 g.ml−1),
S3 (8 × 10−3 g.ml−1) and S4 (1.2 × 10−2 g.ml−1).

b1469_Ch-08.indd 412 4/8/2013 12:34:59 PM


b1469 Catalysis by Ceria and Related Materials

New Developments in Ceria-Based Mixed Oxide Synthesis 413

Figure 8.11 SEM photographs of gadolinium-doped ceria powders derived using


different amounts of polyvinylpyrrolidone surfactant: (a) S1 (2 × 10−3 g.ml−1). (b) S2
(4 × 10−3 g.ml−1). (c) S3 (8 × 10−3 g.ml−1). (d) S4 (1.2 × 10−2 g.ml−1). Reprinted from
Youkun et al.52 with permission from Elsevier.

Trinchi et al.53 synthesized CeO2–TiO2 thin films with this method


for oxygen-gas sensing applications. Yuan et al.54,55 produced a highly
ordered mesoporous Ce1-xZrxO2 solid solution with 2D hexagonal
mesostructures. The synthetic approach was based on a sol-gel pro-
cess combined with evaporation-induced self-assembly in ethanol
using Pluronic P123 as the template and ceric nitrate and zirconium
oxychloride as the precursors without any additional acid or base,
under optimized temperature and humidity conditions. This novel
assembly created structured ceria–zirconia solid solutions having a
high specific surface area and uniform channels, which are good
candidates for applications in catalysis. Typical synthesis steps are
shown in Fig. 8.12. On the whole the modified sol-gel approach has
attracted much attention and several reports can be found in the
literature on the synthesis of well-defined ceria-based nano-oxides.

b1469_Ch-08.indd 413 4/8/2013 12:35:00 PM


b1469 Catalysis by Ceria and Related Materials

414 B. M. Reddy, T. Vinod Kumar and N. Durgasri

Figure 8.12 Assembly of metal precursors with triblock copolymer P123 as struc-
ture-directing agent to form highly ordered mesostructures. Reprinted from Yuan
et al.55 with permission from Elsevier.

8.2.5 Microemulsions
Among the various low-temperature techniques, microemulsion
has received considerable interest for the synthesis of ceria-based
nano-oxides. Microemulsions are colloidal ‘nano-dispersions’ of
water in oil (or oil in water), which are immiscible because of the
very high interfacial energy that exists between the oil/water or
water/oil interface. However, surfactants and cosurfactants reduce
the interfacial energy resulting in the spontaneous mixing of one
liquid in the other. At low concentrations, the surfactant (emulsi-
fier) dissolves in the aqueous phase, but when the concentration
exceeds a critical micellar concentration, the surfactant molecules
organize spontaneously to form aggregates such as micelles, vesi-
cles, etc., because at this concentration the bulk properties of the
surfactant, such as surface tension, conductivity and solubilisation,
change.56,57

b1469_Ch-08.indd 414 4/8/2013 12:35:00 PM


b1469 Catalysis by Ceria and Related Materials

New Developments in Ceria-Based Mixed Oxide Synthesis 415

The choice of surfactant is critical to the size, shape and stability


of the particles. It should be chemically inert with respect to all
other components of the microemulsion. This is particularly impor-
tant when the system contains oxidizing or reducing agents.
Surfactants are classified as cationic, anionic, non-ionic or zwitteri-
onic, depending on the type of charge on their head group. Two of
the most commonly used surfactants are the anionic surfactant
sodium bis (2-ethylhexyl) sulfosuccinate (AOT) and the cationic
surfactant CTAB.
This technique generally uses water-in-oil microemulsions
(reverse micelle). These reverse micelles (referred to as nanoreac-
tors) favour the formation of small crystallites with a narrow size
distribution. On mixing the microemulsion, reverse micelles con-
taining the reactants collide with each other forming a water chan-
nel, which results in the formation of a transient dimer. When the
dimer forms, an intermicellar exchange of reactants takes place and
thus nucleation starts at the micellar edges with a well-known growth
process from the boundary to the core.13,58,59 This method is useful
for controlling the size and morphology of fine particles such as
carbonates, metals, mixed oxides, oxalates, etc.60 Parameters that
effect the microemulsion method are: (i) the water-to-surfactant
ratio, (ii) the solvent effect, (iii) surfactants and cosurfactants, (iv)
surfactant concentration, (v) the nature of the precipitating agent
(reducing agent), etc.59,61
Laguna et al.62 synthesized Fe-doped ceria solids with the micro-
emulsion method. In this process Ce(NO3)3 and Fe(NO3)2 were
used as the precursors, n-octane as a continuous oil phase, CTAB
and 1-butanol as surfactant and cosurfactant and ammonia as pre-
cipitating agent, respectively. Bai et al.63 synthesized comet-like and
prism-like CeO2 nanostructures by the thermal decomposition of a
cerium hydroxycarbonate precursor. These precursors were synthe-
sized under solvothermal conditions at different temperatures. In
this method cerium nitrate was used as the ceria source, urea as the
precipitating agent and ethoxylated nonylphenol (TX-10) and
1-butanol as surfactant and cosurfactant, respectively. The reaction
temperature is crucial when using urea as a precipitant because the

b1469_Ch-08.indd 415 4/8/2013 12:35:00 PM


b1469 Catalysis by Ceria and Related Materials

416 B. M. Reddy, T. Vinod Kumar and N. Durgasri

Figure 8.13 SEM micrographs of CeO2 obtained by microemulsion at a tempera-


ture of 80 °C and 180 °C, and calcination at 450 °C. Reprinted from Jingyi et al.63
with permission from Elsevier.

concentration of OH− ions is directly controlled by the hydrolysis


rate of urea. At low temperatures, urea hydrolyzes very slowly and
the reaction time is too long, and at very high temperatures urea
hydrolyzes very quickly so the reaction time is too short. In both
cases the nucleation rate is different and as a result products with
different morphologies are obtained. SEM images of the resulting
CeO2 calcined at different temperatures are shown in Fig. 8.13. The
use of expensive surfactants is one of the major disadvantages associ-
ated with this method.

8.2.6 Microwave method


Among the various green technologies, microwave chemistry is of
significant importance. Microwaves have both electric and magnetic
field components. They operate in a frequency range between

b1469_Ch-08.indd 416 4/8/2013 12:35:00 PM


b1469 Catalysis by Ceria and Related Materials

New Developments in Ceria-Based Mixed Oxide Synthesis 417

0.3 and 30 GHz. However, for laboratory purposes, a frequency of


2.45 GHz is preferred. Chemical reactions with microwaves are safer,
faster, cleaner and more economical than conventional methods.
Due to these advantages, microwave chemistry has become a full-
grown and useful technique for various applications in recent
years.64–66 It is interesting to note that microwaves neither influence
the orientation of collisions nor the activation energy; the activation
energy remains constant for any particular reaction. However, the
microwave energy affects the temperature parameter. An increase in
the temperature causes more movement of the molecules, which
leads to a larger number of energetic collisions. A plausible reason
for an increase of the effective collision rate under microwave irra-
diation is due to an increase in the collision cross section of the
particles. The enhancement of reaction rates is largely dependent
on many complex factors including vessel size, volume of the pre-
cursor, microwave power delivery and microwave cavity design. It is
quite possible to combine microwave chemistry with the full spec-
trum of available liquid-phase synthesis techniques namely, hydro-
lytic sol-gel, non-hydrolytic sol-gel, hydrothermal, solvothermal,
sonochemistry, biomimetic approaches, co-precipitation and so on.
The broad variety of possible combinations opens up unique and
exciting opportunities for the preparation of inorganic nanoparti-
cles and nanostructures.67–70
Reddy et al.71 synthesized nanosized monophasic Ce0.5Zr0.5O2
solid solutions with a 1:1 mole ratio of component oxides by the
microwave-assisted combustion method. Interestingly, the micro-
wave synthesized sample exhibited better catalytic activity for CO
oxidation than an equivalent sample obtained by conventional co-
precipitation. Raman spectroscopy studies, shown in Fig. 8.14,
revealed that the sample subjected to microwaves contained more
oxygen vacancies with a defect structure cubic fluorite type phase.71,72
Federica et al.73 synthesized Ce0.9Pr0.1O2 nanostructured powders
using microwave energy. In this approach, the Ce0.9Pr0.1O2 powders
were obtained by co-precipitation and the microwaves were applied
for drying and calcination. SEM images of the microwave and con-
ventionally synthesized samples are shown in Fig. 8.15.

b1469_Ch-08.indd 417 4/8/2013 12:35:00 PM


b1469 Catalysis by Ceria and Related Materials

418 B. M. Reddy, T. Vinod Kumar and N. Durgasri

Figure 8.14 Raman spectra of CexZr1−xO2 samples. MWCZ: ceria–zirconia pre-


pared by microwave method. CPCZ: ceria–zirconia prepared by co-precipitation.
Reprinted from Reddy et al.71 with permission from Springer.

The application of microwaves in both drying and calcination


processes of coprecipitated powders seems to be an efficient method
of improving powder quality, leading to shorter processing sched-
ules and enhanced colour development.73 Bondioli et al.74 have syn-
thesized praseodymium-doped ceria powders by a microwave-assisted
hydrothermal method. The powders obtained had nanosized fac-
eted polyhedral morphologies and the particle size was ~30 nm as
shown in Fig. 8.16.

b1469_Ch-08.indd 418 4/8/2013 12:35:00 PM


b1469 Catalysis by Ceria and Related Materials

New Developments in Ceria-Based Mixed Oxide Synthesis 419

(a) (b)

Figure 8.15 SEM micrographs of 10% praseodymia/ceria samples treated at


1000 °C for 15 min. (a) Microwave (10,000×) and (b) traditionally (3,500×)
calcined powders. Reprinted from Bondioli et al.73 with permission from the Royal
Society of Chemistry.

Figure 8.16 TEM micrograph of a 4 mol% praseodymium-doped ceria sample


(14 atm, 120 min). Reprinted from Bondioli et al.74 with permission from the Royal
Society of Chemistry.

b1469_Ch-08.indd 419 4/8/2013 12:35:01 PM


b1469 Catalysis by Ceria and Related Materials

420 B. M. Reddy, T. Vinod Kumar and N. Durgasri

8.2.7 Sonochemical method


The possibility of using sound energy in chemistry was established
more than 70 years ago. By definition, sonochemistry is the applica-
tion of powerful ultrasound radiation (10 kHz to 20 kHz) to cause
chemical changes to molecules. The physical phenomenon behind
this process is acoustic cavitation. Typical processes that occur in
sonochemistry are the creation, growth and collapse of a bubble.75–78
A typical laboratory setup for sonochemical reactions is shown in
Fig. 8.17. More details of sonochemistry and the theory behind it
can be found elsewhere.79,80
Yu et al.81 were the first to prepare nanoporous ceria and ceria–
zirconia solid solutions with high surface area by high-intensity ultra-
sound without thermal post-treatment. In this process they used
(NH4)2Ce(NO3)6 and ZrOCl2·8H2O as metal oxide precursors. The
precursors and urea were added to distilled water and the solution
was sonicated continuously for 3 h, raising the temperature to
around 80°C. Due to the high-intensity ultrasound, collisions
between metal particles take place. These collisions generate local-
ized high-temperature regions, which enhance the condensation
reactions among hydroxyl groups on adjacent hydrous ceria or

Figure 8.17 Typical laboratory-scale sonochemical apparatus. Reprinted with per-


mission from Bang and Suslick.78 Copyright 2010 Wiley-VCH.

b1469_Ch-08.indd 420 4/8/2013 12:35:01 PM


b1469 Catalysis by Ceria and Related Materials

New Developments in Ceria-Based Mixed Oxide Synthesis 421

ceria–zirconia nanoparticles to produce agglomerates of ceria and


ceria–zirconia solid solutions as well as simultaneous crystallization.
Conventional thermal post-treatment at high temperatures is not
required in this method for the crystallization of solid solutions.81
Guo et al.82 synthesized ceria–zirconia mixed oxides using polyethyl-
ene glycol (PEG) with the sonochemical method. Zhang et al.83 pro-
duced CeO2 nanorods (<10 nm diameter) using the PEG surfactant
as the structure-directing agent via ultrasonication at room tempera-
ture. To remove the PEG, the sample was washed repeatedly with
alcohol and subsequently dried at 100°C in air for 1 day. Their study
revealed the significance of ultrasonication, without which it was dif-
ficult to obtain nanorods. TEM images, shown in Fig. 8.18,

Figure 8.18 TEM images of ceria nanorods prepared by the sonochemical method.
Reprinted with permission from Zhang et al.83 Copyright 2007 the American
Chemical Society.

b1469_Ch-08.indd 421 4/8/2013 12:35:01 PM


b1469 Catalysis by Ceria and Related Materials

422 B. M. Reddy, T. Vinod Kumar and N. Durgasri

confirmed that with the increase of sonication time, more and more
nanoparticles disappeared and nanorods were the main product.
Sonochemical reaction rates depend on the irradiation fre-
quency, acoustic power and volatility of the solvent, which influence
the cavitation strength. From a preparation point of view, sono-
chemical approaches are advantageous because they speed up reac-
tion rates and favour the formation of porous products with a high
specific surface, and have the additional economic benefit of rela-
tively low reaction temperatures.83,84

8.2.8 Combustion synthesis


Of the various synthetic methods, combustion synthesis (CS) has
special significance because CS processes are characterized by high
temperatures, fast heating rates and short reaction times. These
features make CS an attractive method for the manufacture of mate-
rials. In this method, the exothermicity of the redox (reduction–
oxidation or electron transfer) chemical reaction is used to produce
useful materials.85,86
The important events that occur during combustion are the
generation of heat and gas evolution. The heat generated helps in
the crystallization and the formation of the desired phase. On the
other hand, a very high flame temperature adversely affects crystal-
lite size due to agglomeration and thereby a reduction in surface
area. Sintering is expected but the evolution of gases during com-
bustion dissipates the heat of combustion and limits the rise in tem-
perature, thus reducing the possibility of premature local partial
sintering of the primary particles. The powder characteristics are
primarily governed by the heat of combustion and gas evolution,
which themselves are dependent on the nature of the fuel and the
oxidant-to-fuel ratio. The commonly used fuels in this process are
glycine, citric acid, urea, ascorbic acid, etc.87,88
Different types of combustion techniques are employed for the
synthesis of nanomaterials. These methods are classified on the
basis of the physical nature of the reaction media as: (i) flame syn-
thesis or gas phase combustion, (ii) solution combustion synthesis

b1469_Ch-08.indd 422 4/8/2013 12:35:01 PM


b1469 Catalysis by Ceria and Related Materials

New Developments in Ceria-Based Mixed Oxide Synthesis 423

Figure 8.19 Solution combustion process before, during and after the reaction.90
Reprinted from Mukasyan et al.90 with permission from Springer.

where the initial reaction media is an aqueous solution and


(iii) heterogeneous condensed phase (solid) combustion.89 Of
these, solution combustion synthesis is attractive because the initial
reaction media are in the liquid state (e.g., an aqueous solution),
which allows mixing of the reactants on a molecular level, thus
permitting precise and uniform formulation of the desired compo-
sition on the nanoscale. Typical steps in this process are shown in
Fig. 8.19. Depending on the type of procedure, combustion synthe-
sis in the liquid phase is again broadly divided into various types:
gel combustion, sol-gel combustion, emulsion combustion, thermal
explosion, etc.86,90
The synthesis of fine particles of ceria by the combustion
method has been investigated since 1990.91,92 Chinarro et al.93 synthe-
sized CeO2–Gd2O3–CaO and CeO2–Gd2O3–Y2O3 ternary oxides
using urea as the fuel and Ce(NO3)3·6H2O, Ca(NO3)2·4H2O,
Gd(NO3)3·6H2O and Y(NO3)3·6H2O as cation precursors. The stoi-
chiometric composition of each mixture was calculated based on
total oxidising and reducing valences of the oxidizer and the fuel, in
order to release the maximum energy for the reaction. SEM images
of powders prepared using this procedure are shown in Fig. 8.20
Baidya et al.94 exploited this method to prepare nanosized
Ce0.73Ti0.25Pd0.02O2−δ solid solutions for CO oxidation, where glycine
was used as the fuel and (NH4)2Ce(NO3)6, PdCl2 and TiCl4 were the
metal sources. The TiCl4 was first converted into TiO(NO3)2 in situ
by using water and HNO3. After synthesis, repeated washing
removed the chloride ions.

b1469_Ch-08.indd 423 4/8/2013 12:35:01 PM


b1469 Catalysis by Ceria and Related Materials

424 B. M. Reddy, T. Vinod Kumar and N. Durgasri

Figure 8.20 SEM micrographs of the as-prepared powder of (a) CeO2–Gd2O3–


CaO and (b) CeO2–Gd2O3–Y2O3. Reprinted from Chinarro et al.93 with permission
from Elsevier.

Supported catalysts for various catalytic applications are nor-


mally prepared using the so-called impregnation method. In gen-
eral, the porous solid support is saturated with the desired catalytic
agent. Many techniques are available to facilitate the impregnation
process: pressurization, vacuum treatment, acoustic activation, etc.
However, most of them are time consuming and expensive. Recently,
a novel method for the synthesis of such materials in one step using
a controllable self-propagating combustion mode, impregnated
inert support combustion (IISC), was developed.95,96 However, this
method is not applicable for all types of material. Low exothermic
mixtures do not have a self-sustained mode of reaction propagation
during impregnation on the inert support media. To overcome this
obstacle, another method, impregnated active layer combustion
(IALC), was developed.96

8.2.9 Impregnation method


To render a metal/salt component into a finely divided form on a
support requires a dispersion stage achieved by impregnation.
Impregnation for preparing a supported catalyst is achieved by

b1469_Ch-08.indd 424 4/8/2013 12:35:02 PM


b1469 Catalysis by Ceria and Related Materials

New Developments in Ceria-Based Mixed Oxide Synthesis 425

filling the pores of a support with a solution of the metal salt from
which the solvent is subsequently evaporated. The impregnation
method has two main steps. The first step is to deposit the active
component precursor, mostly using aqueous solutions, onto a sup-
port, and the second step is to transform this precursor into the final
active phase, which often involves drying and heat treatment (calci-
nation or reduction). In this process the metal compounds are held
by adsorption of metal cations on a basic site of the support (O2−,
OH−).3,60,97
Zhao et al.98 used a low-temperature polymerizable metal com-
plex solution as an active component precursor to prepare sup-
ported (structured) catalysts by a sequence of impregnation, drying
and calcination. The polymerizable metal complex method is often
referred to as the Pechini method.99 In this method polymerization
between citric acid and ethylene glycol or polyethylene glycol takes
place. The strongly chelating ability of citric acid with most metal
ions frequently facilitates the formation of a metal citrate complex
during polymerization. A 5 wt% ZrxCe1-xO2/Al2O3 and structured
carbon nanofibre (SCNF) supported Cu–CeO2 were prepared using
this method. Amounts of ZrO(NO3)2.xH2O, Ce(NO3)3.6H2O and
citric acid (C6H8O7) in a weight ratio of 1:1:2 were added to deion-
ized water to form the complex solution. PEG was added and then
a weighed amount of alumina powder was gradually added under
vigorous stirring to produce a suspension. The suspension was sub-
sequently evaporated to form the dried solid. A similar procedure
was carried out to prepare SCNF supported Cu–CeO2, which is sche-
matically shown in Fig. 8.21. Forming a stable active component
complex with a suitable support is a promising strategy for avoiding
the drawbacks of the impregnation method.
Tang et al.100 prepared a CeO2–CoOx catalyst for CO oxidation by
using impregnation, precipitation-oxidation, co-precipitation and
the hydrothermal method. Of these methods, the impregnation
method yielded well-dispersed ceria strongly interacting with the
cobalt oxide. The impregnated catalyst was the most active for CO
oxidation followed by precipitation-oxidation, co-precipitation and
the hydrothermal method.100 Several publications can be found in

b1469_Ch-08.indd 425 4/8/2013 12:35:02 PM


b1469 Catalysis by Ceria and Related Materials

426 B. M. Reddy, T. Vinod Kumar and N. Durgasri

Figure 8.21 Formation of supported catalysts using complexes.98 Reprinted from


Zhao et al.98 with permission from Springer.

the literature on various supported catalysts prepared by the impreg-


nation method.

8.2.10 Electrochemical method


The electrochemical method is an exciting phenomenon. In this
process, materials can be prepared by passing an electric current
through an electrochemical cell from an external source. The
important features that make this method well suited for nanotech-
nology, biotechnology and microtechnology are: (i) the experi-
ments are simple to perform and the instruments required are
inexpensive and readily available, (ii) electrochemical synthesis
takes place at the electrode/electrolyte interface, (iii) the product is
normally deposited on the electrode in the form of a thin film or a
coating, (iv) electrochemical synthesis is a low-temperature process
limited by the boiling point of the electrolyte and (v) it can be per-
formed near room temperature from water-based electrolytes.
Reactions can be controlled kinetically by controlling the current in
the cell, and they can be thermodynamically controlled by choosing
the applied potential.
The two important parameters that determine the course of the
reaction are: (i) the deposition current and (ii) the cell potential.

b1469_Ch-08.indd 426 4/8/2013 12:35:02 PM


b1469 Catalysis by Ceria and Related Materials

New Developments in Ceria-Based Mixed Oxide Synthesis 427

Figure 8.22 (a) SEM, (b) TEM, high-resolution transmission electron microscopy
(inset), and selected area electron diffraction (inset) images of Ce1−xCoxO2−δ
nanorods. Reprinted with permission from Ou et al.103 Copyright 2010 the American
Chemical Society.

Of the two, either can be controlled as a function of time during the


reaction. The morphology and composition of the product depends
on the current density, the nature of anions or cations in the solu-
tion, the bath composition, temperature, the solution concentra-
tion, the power supply current waveform, the physicochemical
nature of the substrate surface, etc.101,102
Ou et al.103 exploited this method to make Ce1−xCoxO2−δ nanorods
at room temperature as shown in Fig 8.22. The process was carried
out in a solution of Ce(NO3)3, Co(NO3)2 and EDTA disodium salt
with a current density of 4.0 mA.cm−2 for 60 min at room tempera-
ture. A simple three-electrode cell was used for this experiment.
During the process, the oxidation state of cerium changed from 3+
to 4+ in the alkali solution. During the electrochemical deposition,
NO3− ions in the deposition solution were first electro-reduced to
form OH− ions on the surface of the cathode, Eq. (8.1). These OH−
ions are responsible for the formation of CeO2 and Co3O4 via
Eqs. (8.2) and (8.3), respectively. It is well known that electrochemi-
cal deposition mixes the chemicals at an atomic level. Therefore,
mixed CeO2 and Co3O4 at an atomic level are always obtained during

b1469_Ch-08.indd 427 4/8/2013 12:35:02 PM


b1469 Catalysis by Ceria and Related Materials

428 B. M. Reddy, T. Vinod Kumar and N. Durgasri

electrochemical deposition. Finally homogeneous Ce1−xCoxO2−δ


composites were obtained via Eq. (8.4).

NO3− + H2O + 2e → NO2− + 2OH− (8.1)


4Ce3+ + 12OH− + O2 → 4CeO2 + 6H2O (8.2)
6Co2+ + 12OH− + O2 → 2Co3O4 + 6H2O (8.3)
3(1− x)CeO2 + xCo3O4 → 3Ce1−xCoxO2−δ (8.4)

The same group also produced Ce1-xZrxO2 flower-like nanostruc-


tures, hierarchically porous Gd3+-doped CeO2 nanostructures, etc.,
with the same electrodeposition method.104,105 There is much scope
for making well-designed catalysts through this method.

8.2.11 Ball milling


Most of the preparation methods described earlier involve and
depend mainly on a nucleation step. Mechanochemical methods
are entirely different from chemical methods. The ball-milling tech-
nique is a more environmentally safe method than chemical synthe-
sis, producing far less chemical waste without the need for large
amounts of solvents.106–108 Ball milling is a physical method, which is
a top-down method. It is an efficient tool for grinding various mate-
rials into fine powders. There are two types of grinding processes:
wet grinding and dry grinding. The production of nanoparticles
from dry grinding is difficult because agglomeration takes place
when they contact each other. Accordingly, to minimize agglomera-
tion grinding has to be carried out in a liquid (wet grinding). In this
approach, a solvent surrounds the surface of the generated particle
instantaneously and it is possible to control the agglomeration of
particles much better than in dry milling.109
The primary objective of milling is often purely particle size
reduction. Milling produces fine particles by applying mechanical
energy to solid materials to break the bonding between atoms or
molecules. In this process powder particles are trapped between
highly kinetic colliding balls and the inner surface of the vial, which
causes repeated deformation, rewelding and fragmentation of

b1469_Ch-08.indd 428 4/8/2013 12:35:02 PM


b1469 Catalysis by Ceria and Related Materials

New Developments in Ceria-Based Mixed Oxide Synthesis 429

Figure 8.23 Changes in a material trapped between colliding balls in a ball mill.110

premixed powders resulting in the formation of finely dispersed


particles in the grain-refined matrix.110 This process is schematically
shown in Fig. 8.23.
Different types of ball-milling techniques are available for parti-
cle size reduction based on the movement of milling balls and vial,
such as the planetary mill, the vibration mill and the attritor. In ball
milling, particle size reduction mainly depends on the rotation
speed, size of the balls, weight ratio of balls to powder, medium of
milling, milling time, etc. Because of the variety of powder materials,
the selection of parameters varies significantly. Contamination of
the material can arise from several sources, which include: (i) vials
and grinding media, (ii) milling atmosphere, (iii) control agents
added to the powders and (iv) impurities in the starting powders.
The extent of contamination increases with increasing milling
energy and milling time. One way of minimizing contamination
from the grinding medium and the container is to make them out
of the same materials as the powder being milled.111
Lia et al.112 prepared nanosized ceria by a wet solid phase mecha-
nochemical method using Ce2(CO3)3.8H2O as ceria source with
sodium hydroxide and some amount of water. The mechanochemi-
cal reaction process was divided into two steps. The first step was a
multi-phase mechanochemical reaction, where the crystalline

b1469_Ch-08.indd 429 4/8/2013 12:35:02 PM


b1469 Catalysis by Ceria and Related Materials

430 B. M. Reddy, T. Vinod Kumar and N. Durgasri

Ce2(CO3)3.8H2O was trapped between colliding balls or between a


ball and the vial, and subjected to a severe plastic deformation. In
the next step the material was reacted with NaOH to form amor-
phous Ce(OH)4. There is a possibility of Ce(OH)3 formation but the
formation free energy (−∆G) of Ce(OH)4 is more than that of
Ce(OH)3. The final step is the crystallization of Ce(OH)4. The reac-
tion and crystallization rates are enhanced with the presence of
water, which provides a suitable route for the material transforma-
tion. The thermodynamic driving force is quite favourable for CeO2
formation at room temperature (∆G = −263 kJ). However, the kinet-
ics of the solid-state reaction is extremely slow.
Various ceria-based mixed oxides have been prepared by ball
milling, including samarium-doped cerium oxide,113 CeO2–ZrO2,114
Ce0.8Gd0.2O2−δ,115 etc.

8.2.12 Spray pyrolysis


In recent years, spray-pyrolysis synthesis has received particular
attention for the synthesis of nanomaterials and is widely used for
the preparation of photocatalysts, phosphors, Li-ion batteries,
porous materials, materials with metastable phases, etc. Spray-
pyrolysis thin-film synthesis techniques produce nanocrystalline
electrolyte thin films for solid oxide fuel cells with average grain
sizes below 100 nm and film thicknesses between 100 and 500 nm.116
This technology is an inexpensive, continuous and ambient pressure
process. The technique was first developed by the research group of
Sotiris E. Pratsinis at ETH Zurich, Switzerland.117 It is a promising
aerosol process for producing designer particles of precisely con-
trolled morphology with decorations on the surfaces or inside the
particles. In principle, it is similar to combustion synthesis, in that
liquid fuel is oxidized and produces gases and particles. The only
difference is that spray pyrolysis produces useful powders, whereas
combustion generates pollutant particles.118,119
As per the literature, there are two types of approach: vapour
feed-flame spray pyrolysis and liquid feed-flame spray pyrolysis
(LF-FSP). However, it is often difficult in the vapour feed process to

b1469_Ch-08.indd 430 4/8/2013 12:35:02 PM


b1469 Catalysis by Ceria and Related Materials

New Developments in Ceria-Based Mixed Oxide Synthesis 431

produce multicomponent materials with homogeneous composi-


tion. LF-FSP has the potential to make a wide variety of single and
mixed-metal oxide nanopowders in a single step. It provides easy
and precise compositional control and is a low-cost route to versatile
materials in one-step processing. This method easily produces many
types of nanopowder with excellent control of phase purity and mor-
phology. To date LF-FSP has been used to make phase-pure nano-
powders of single and mixed-metal oxides including CeO2, Al2O3,
ZrO2, TiO2, CoO and NiO.120
During spray-pyrolysis deposition, a precursor solution of metal
salt and solvent is sprayed as fine droplets onto a heated substrate.
When the droplets reach the heated substrate, they spread out and
undergo pyrolytic decomposition. Water evaporates from the drop-
lets and the precursor is converted to an oxide with the formation
of solid particles.121–123
Rupp et al.123 synthesized nanocrystalline gadolinium-doped
ceria electrolytes (CGO) of 100–500 nm thickness. These thin films
had advantages compared to sintered electrolyte pellets or tapes.
Pichestapong et al.122 prepared samarium-doped ceria nanoparticles
with 0.1 M starting solutions of cerium nitrate containing samarium
nitrate between 0 and 20 mol% using a pyrolyzing temperature of
600 °C. Their study suggests that one oxide particle is formed from
one droplet of the precursor. More recently, CeO2-doped TiO2
nanopowders composed of single-crystalline spherical particles with
primary particle size of 10–13 nm and Ce doping concentrations of
5–50 at% have been synthesized.124 Interestingly, square-shaped
particles with an average size of around 8.2 nm have been observed
for flame-made CeO2.

8.3 Catalytic Combustion and Oxidation Applications


During the last century concerns over the availability of energy
sources and the impact of combustion processes on the global envi-
ronment have attracted more and more attention. The imperative
reasons for the use of catalysts in combustion may be either to gen-
erate energy in catalytic heaters, boilers and gas turbine engines, or

b1469_Ch-08.indd 431 4/8/2013 12:35:03 PM


b1469 Catalysis by Ceria and Related Materials

432 B. M. Reddy, T. Vinod Kumar and N. Durgasri

to remove or control the formation of pollutants (volatile organic


compounds or VOCs, CO, soot, etc.) from exhausts or industrial
units.125,126 Furthermore, concerns about atmospheric pollution
have been continuously growing since the mid-1950s, when the
harmful effects of emissions from combustion sources were identi-
fied. In contrast to conventional combustion, catalytic combustion is
generally defined as complete oxidation promoted at the surface of
a catalyst; it is a flameless process. In this process, the selection of the
catalytic material is important. Among many materials, ceria-based
materials are of significant importance in combustion applications
due to their special features.

8.3.1 Catalytic combustion of VOCs


VOCs are one of the hazardous pollutants found in air. They are
organic compounds with boiling points in the range 50–260 °C at
room temperature having an atmospheric pressure of 133.322 Pa
and are widespread in the atmosphere. The sources of VOCs can be
either indoors (from household products such as office supplies,
insulating materials, cleaning products and pressed woods) or out-
doors (mainly industrial processes and automobile exhausts). The
emission of VOCs has received particular attention since they are
involved in photochemical smog, the depletion of atmospheric
ozone and the production of ground-level ozone. VOCs generally
have mutagenic, carcinogenic or teratogenic effects on humans and
so environmental legislation has imposed severely stringent targets
for the permitted levels of VOC emissions.127,128
Physical adsorption and thermal incineration have been devel-
oped for the abatement of industrial VOC emissions in recent dec-
ades. Physical adsorption is an effective way of eliminating VOCs
with porous materials for a short period, but is not promising due to
the limited removal capacity. Thermal incineration is a convenient
way to remove VOCs. However, this requires a rather high tempera-
ture (usually above 800–1200 °C), consumes a large amount of
energy and generates noxious by-products like NOx. Therefore, one
of the most effective and economically practicable VOC removal

b1469_Ch-08.indd 432 4/8/2013 12:35:03 PM


b1469 Catalysis by Ceria and Related Materials

New Developments in Ceria-Based Mixed Oxide Synthesis 433

technologies is catalytic combustion. It is a flameless process and can


operate over a temperature window of 500–1000 °C. Interest in this
process has been explored for two reasons: for pollution abatement
since it is an environmentally benign technology and for the pro-
duction of heat and energy (utilized in catalytic heaters, boilers and
gas turbine engines). Catalytic combustion can take place with
leaner air-to-fuel ratios than required for flame combustion. Another
advantage is that the combustion produces fewer harmful by-prod-
ucts since the surface-catalyzed free-radical chain mechanism occurs
in combustion. Thus, the ultimate results are economic and envi-
ronmental benefits.129,130 A simple catalytic combustion setup is
shown in Fig. 8.24.
Although catalytic total oxidation reactions have been studied
since 1817, the earliest work on catalytic combustors was first
described by Pfefferle in 1917. Active interest in catalytic combus-
tion for power generation increased during the early 1990s. For
practical applications of catalytic combustion, the light-off tempera-
ture is a critical parameter that determines both system operation
and system design as suggested by Pfefferle.131,132
Even though many catalytic materials have been evaluated for
combustion applications, ceria-based materials have received par-
ticular interest in this regard. For efficient combustion the choice
of catalyst depends on its physical form and performance over a
wide range of operating conditions, and also on the VOC molecules
to be treated. A multiple approach for combustion involves noble
metals and their support materials to provide environmental

Figure 8.24 Precision combustion two-stage catalytic combustion system.131

b1469_Ch-08.indd 433 4/8/2013 12:35:03 PM


b1469 Catalysis by Ceria and Related Materials

434 B. M. Reddy, T. Vinod Kumar and N. Durgasri

protection against VOCs. Amongst the various noble metals, Pd and


Pt have received particular attention for lower alkane combustion.
A wide variety of problems in energy conversion and pollution con-
trol are likely to be solved using metal-supported platinum-catalyst
combustors.129
Noble-metal-based catalysts are known to be active for hydrocar-
bon combustion at low temperatures but are expensive and sinter
easily. For a mixture of several types of hydrocarbon, as in diesel or
jet fuel, a combination of several different metals could be used.
Paucity and economic considerations, low-energy consumption and
the high volatility of noble metals have necessitated the search for
alternative low-cost materials for combustion applications. Hence,
the better properties of metal oxides are promising options against
simple metals. Numerous experimental data available in the litera-
ture indicate that metal-oxide-based catalysts with a reasonably large
specific surface area have good combustion activity, which seems to
correlate with the strength of the metal–oxygen bond.
The use of ceria and ceria-based materials in catalytic science
has been well established in recent decades. Several investigations
have reported that different kinds of interaction between ceria and
various metal oxides influence catalytic activity and facilitate an
improvement in the dynamic performance in the removal of VOCs.
Catalytic activity is a complex function of several factors, such as
oxidation state, nature of the support, catalyst composition, prepa-
ration method, crystal structure, particle size and so on. Doping of
ceria creates an additional structural defect that influences the oxy-
gen storage/release properties. Further, oxygen ion mobility affects
phase stability and the aging properties of CeO2, and increases cata-
lytic activity.133

8.3.1.1 Ceria in combination with zirconia


Among many catalytic systems, CeO2–ZrO2 (CZ) has a special tech-
nological importance and its importance has been substantiated in
the catalytic combustion of several hydrocarbons. CZ mixed oxides
combine acid sites along with easily accessible lattice oxygen

b1469_Ch-08.indd 434 4/8/2013 12:35:03 PM


b1469 Catalysis by Ceria and Related Materials

New Developments in Ceria-Based Mixed Oxide Synthesis 435

species.133 It has been observed that the incorporation of zirconium


cations into the ceria lattice noticeably lowers light-off temperatures
in the combustion of chlorinated molecules. Rivas et al.134 reported
that combustion of dichloroethane (DCE) takes place at a signifi-
cantly lower temperature over a CZ catalyst. The primary combus-
tion products were CO2, HCl and Cl2. It was also proved that the
specific surface area of alumina containing a CZ (Ce0.8Zr0.2Al0.4)
catalyst is three times larger than that of Ce0.8Zr0.2O2, because the
Al2O3-coating increases the surface area of the calcined and aged
materials, it enhances their OSC, and thereby increases combustion
activity.133
The activity of CZ and CZ supported by CuO in the combustion
of ethyl acetate was evaluated by Yang et al.135 The catalytic activity of
these two samples was lower than that of CuO/CZ/TiO2, which
exhibited an enhanced catalytic activity and CO2 selectivity. The
reducibility of surface and mobility of bulk oxygen species in Co3O4/
Z composite oxides correlated with CH4 combustion activity.136 It has
been observed that a Pt/Ce0.67Zr0.33O2 catalyst is more active for CH4
combustion; its activity is even higher than a Pt/Al2O3 catalyst.137
As reported by Wang et al.,138 the combustion of ethanol over a
single Ce0.5Zr0.5O2 catalyst is quite low, T50 was 197 °C. But catalytic
activity was greatly promoted after loading with Pt. For Pt/γ–Al2O3,
T50 was 163 °C. However, for a Pt/γ–Al2O3/CeO2 catalyst, T50 was 153
°C. Interestingly, a CexZr1−xO2 mixed oxide exhibited a significant
reduction in T50 from 10 to 30 °C with an increase in the loading of
zirconia in the ceria lattice. The order of catalytic activity over cata-
lysts is as follows: Pt/γ–Al2O3/Ce0.5Zr0.5O2 (127 °C) > Pt/γ–Al2O3/
Ce0.25Zr0.75O2 (142 °C) > Pt/γ–Al2O3/Ce0.75Zr0.25O2 (146 °C) > Pt/
γ–Al2O3/CeO2 (153 °C) > Pt/γ–Al2O3/ ZrO2 (158 °C).
By loading low-surface-area perovskite species on high-surface-
area active supports, activity can be improved. Alifanti et al.139 loaded
10 and 20 wt% LaCoO3 perovskites on Ce1−xZrxO2 (0 < x < 0.3) sup-
ports and investigated the combustion of benzene and toluene. The
ceria–zirconia-supported LaCoO3 catalyst was found to be very active
in the combustion of benzene and toluene in diluted streams. This
catalyst decreased the light-off temperatures and increased the

b1469_Ch-08.indd 435 4/8/2013 12:35:03 PM


b1469 Catalysis by Ceria and Related Materials

436 B. M. Reddy, T. Vinod Kumar and N. Durgasri

reaction rates by an order of magnitude compared to the bulk per-


ovskite catalysts.

8.3.1.2 Ceria with IB group metals


Cu-doped ceria is a good candidate for catalytic combustion. The
incorporation of Cu species into the CeO2 lattice enhances its OSC
and redox property by decreasing the activation energy to much
lower temperatures for the reduction of Ce4+, as revealed by the
investigations of Hu et al.140 Their study discovered that Ce0.87Cu0.13Oy
catalyst exhibits superior activity for the combustion of acetone. It
was also observed that acetone conversion over the fresh catalyst
decreases sharply when increasing the pulse number, which is
expected to be due to the depletion of lattice oxygen in the
catalyst.140,141
More recently, Cu/CeO2 catalysts have been prepared in a differ-
ent way using electrospinning and have been evaluated for methane
combustion by Ontelli et al.142 A high temperature was reached dur-
ing combustion, which modified the surface area of the catalysts
owing to sintering. However, the reduction in the surface area did
not affect the nature of the active sites. They proposed that the best
results are due to the introduction of copper oxide, which increases
the number of defects in the crystal lattice of the catalyst, and
thereby improves catalyst performance.
It could be noted from the literature that the preparation
method has an important influence on the catalytic activity of Au/
CeO2 catalysts, since the preparation method affects the dispersion
of gold on the ceria support. Au/CeO2 catalysts exhibited better
activity in the combustion of toluene than cerium oxide alone,
because in the former case the gold weakens the strength of surface
Ce–O bonds.143 Several investigations showed that gold is more
effective in improving the catalytic behaviour of ceria-based catalysts
when it was added by deposition-precipitation (DP), because it pro-
motes fine particles. Thus Au/CeO2 prepared by DP showed toluene
conversion starting at about 200 °C, and 100% conversion was
achieved at 360 °C.144 In contrast, a catalyst prepared by

b1469_Ch-08.indd 436 4/8/2013 12:35:03 PM


b1469 Catalysis by Ceria and Related Materials

New Developments in Ceria-Based Mixed Oxide Synthesis 437

co-precipitation (CP) showed higher light-off temperatures than the


DP catalyst.139
Centeno et al.145 completely oxidized benzene and n-hexane at
300 °C with a T50 of 265 °C and at 370 °C with a T50 of 305 °C over a
2.5Au/CeO2/Al2O3 catalyst. Also the complete oxidation of 2-pro-
panol to CO2 was noted with a T50 of 280 °C on the same catalyst
showing the favourable influence of CeO2.
As reported by Scirè et al.,146 the activity of ceria supported by IB
group metals prepared by DP for methanol combustion followed
the order: Au/CeO2 > Ag/CeO2 > Cu/CeO2 > CeO2. On the other
hand, for the same catalysts obtained by CP the order of activity was
Ag/CeO2 > Au/CeO2 > Cu/CeO2 > CeO2 (Fig. 8.25). Interestingly,
for all the catalysts the VOC reactivity order was methanol > acetone
> toluene.146 This trend is in accordance with the degree of interac-
tion between the organic molecules and the catalyst surface.

8.3.1.3 Ceria in combination with other transition elements


As reported by Wang et al.147 MnOx–CeO2 is a superior catalyst for
chlorobenzene (CB) combustion. The incorporation of Mn cations
into the ceria lattice facilitates more active oxygen generation and
improves catalytic activity towards CB combustion. Further, the

Figure 8.25 Methanol conversion vs. reaction temperature over IB metal/ceria


catalysts prepared by (A) DP and (B) CP. Reprinted from Scirè et al.146 with permis-
sion from Elsevier.

b1469_Ch-08.indd 437 4/8/2013 12:35:03 PM


b1469 Catalysis by Ceria and Related Materials

438 B. M. Reddy, T. Vinod Kumar and N. Durgasri

removal of chlorine species from the surface of MnOx–CeO2 cata-


lysts is significantly promoted by an increase of Mn content. Among
various samples investigated, the Mn0.86Ce0.14O2 catalyst exhibited
the best activity with complete combustion at 236 °C with a gas
hourly space velocity (GHSV) of 15000 h−1. Adsorption of CB on the
catalyst surface helps C–Cl bond dissociation even at lower tempera-
tures, and the main combustion products observed were HCl, Cl2,
CO2 and trace amounts of CO.147 It should be noted that the catalytic
combustion of VOCs usually requires a high GHSV varying up to
30000 ml.g−1.h−1. Wu et al.148 found that the conversion efficiency
and reaction rate over a Mn0.2Ce0.8/Al2O3 catalyst were maximum
and decreased with an increase of GHSV at 300 °C. This study found
more than 90% conversion and 99.5% selectivity towards carbon
oxides with a GHSV of 7500 h−1.
It is known that ceria does not react readily with aluminium or
boron oxides but Zarur et al.149 were able to deposit up to 25 wt%
ceria on boron hexaaluminate (BHA) nanoparticles without form-
ing a separate ceria domain. The ceria deposited on BHA preserved
the nanocrystalline morphology up to a calcination temperature of
1300 °C. The CeO2–BHA nanoparticles allowed light-off of a stream
of 1 vol% CH4 in air at a low temperature of 400 °C, compared to
540 °C and 690 °C required for Mn- and Co-substituted BHA sys-
tems, and attained full CH4 conversion by 600 °C. In addition, vari-
ous ceria-based materials have been used for VOC combustion with
different light-off temperatures as shown in Table 8.2.

8.3.2 Soot combustion and oxidation


Soot particulates are an important component of air pollution and
are harmful to both the environment and human beings due to
their potential for mutagenic and carcinogenic activity. Therefore,
research into soot abatement in the exhaust from diesel engines is
receiving much interest. Ceria-based materials have attracted con-
siderable attention for the oxidation of diesel soot particles.151 The
use of CeO2-supported metals can enhance oxidation owing to the
interaction between the metal and the support, which improves the

b1469_Ch-08.indd 438 4/8/2013 12:35:03 PM


b1469 Catalysis by Ceria and Related Materials

New Developments in Ceria-Based Mixed Oxide Synthesis 439

Table 8.2 Catalytic combustion of VOCs on various ceria-based catalysts.


Method of BET SA Combusted T50
Catalyst preparation (m2.g−1) material (°C) Reference
Ce0.5Zr0.5O2 CP 99 DCE 295 134
TCE 390
Co3O4/CZ IMP 19 CH4 445 136
20% LaCoO3/CZ Wet IMP 37.9 Toluene 192 139
Pd/CeO2−Al2O3 Sol-gel 203 CH4 < 400 150
Mn0.2Ce0.8O2/Al2O3 IMP 136 CB 361 147
Pt/Al2O3/Ce0.5Zr0.5O2 IMP 99.6 Ethanol 127 138
Au/CeO2/Al2O3 DP 85 Benzene 265 145
n-hexane 370
2-propanol 280
Co3O4/CeO2 CP 28 CH4 400 136
Cu0.13Ce0.87O2 Combustion 27.1 Acetone 200 140
Au/CeO2 DP 50 Toluene 225 144
MCD 46 250
CP 96 270
CP = Co-precipitation, IMP = Impregnation, DP = Deposition precipitation, MCD = Metallic
colloid deposition.

redox characteristics of ceria. The soot oxidation/combustion tem-


perature depends on various factors like soot type and composition,
catalyst material, particle size and size distribution, and the contact
efficiency between the soot and the catalyst. It is also known that
eutectic mixtures and low-melting-point catalysts boast superior per-
formance for complete soot oxidation.152,153
Environmental transmission electron microscopy (ETEM) allows
direct observation of the soot/catalyst interface to monitor in situ
oxidation of the soot. Thermogravimetric (TG) and temperature-
programmed methods have been used to study soot combustion in
the presence and absence of oxygen, and the properties depend on
the accessibility of gaseous oxygen and the location, shape and
dimension of the ceria/soot interface (Fig. 8.26). Instead of using
ceria on its own as a catalyst for soot oxidation, more studies have

b1469_Ch-08.indd 439 4/8/2013 12:35:03 PM


b1469 Catalysis by Ceria and Related Materials

440 B. M. Reddy, T. Vinod Kumar and N. Durgasri

Figure 8.26 TEM image of contact between ceria and soot.154 Reprinted from
Simonsen et al.154 with permission from Elsevier.

focused on the modification of CeO2 with different cations to


increase activity and thermal stability.154
As mentioned earlier, a ceria–zirconia solid solution promotes
bulk oxygen mobility. Different compositions at different calcina-
tion temperatures affect soot combustion, as proposed by Aneggi
et al.155 and as shown in Fig. 8.27.
Soot combustion depends on the nature of the catalyst material
and the type of soot. Atribak et al.156 prepared Ce0.76Zr0.24O2 catalysts
by co-precipitation (CP) and reverse microemulsion (RME) having
BET surface areas of 67 and 128 m2.g−1, respectively. The activity of
a commercial catalyst from Rhodia (RH) (Ce0.75Zr0.25O2, 113 m2.g−1)
was also used in testing. The catalytic activity of the various Ce–Zr
catalysts tested was in the order: RH > RME > CP. This trend can be
explained based on the chemical properties of the catalysts rather

b1469_Ch-08.indd 440 4/8/2013 12:35:03 PM


b1469 Catalysis by Ceria and Related Materials

New Developments in Ceria-Based Mixed Oxide Synthesis 441

Figure 8.27 T50 against surface area for each composition as obtained from TG
experiments over a soot/catalyst mixture under air. Reprinted from Aneggi et al.155
with permission from Elsevier.

than the BET surface area. It was also observed that all Ce–Zr cata-
lysts tested were more effective for soot combustion than a reference
Pt catalyst, due to the active oxygen-assisted mechanism. The extent
to which Ce–Zr catalysts accelerate the combustion of Printex-U (as
well Vulcan XC72R) commercial soot depends on the nature of cata-
lyst, as shown in Fig. 8.28.
It was reported that CeO2–Fe2O3 mixed oxides improve the
kinetic performance of soot combustion. The reaction proceeds
through a redox cycle between Ce4+/Fe3+ and Ce3+/Fe2+, and Fe2O3
has been shown to exhibit a ‘push–pull’ redox mechanism for soot
combustion at higher temperatures. The active sites involved are
Fe–O–Ce-type species. The mechanism shown in Fig. 8.29 involves
the surface oxide anion bound to the Fe3+, which reacts with the soot

b1469_Ch-08.indd 441 4/8/2013 12:35:03 PM


b1469 Catalysis by Ceria and Related Materials

442 B. M. Reddy, T. Vinod Kumar and N. Durgasri

Figure 8.28 Printex-U combustion by NOx+O2, in a tubular reactor coupled to gas


analysers under loose-contact. Reprinted from Atribak et al.156 with permission from
Elsevier. (COP = co-precipitation)

at the soot/catalyst interface to produce COx with the formation of


oxygen vacancies and reduced Fe2+.157
Aneggi et al.158 investigated various silver-doped ceria (1–10 wt%)
catalysts prepared by incipient wetness impregnation. A Ce5Ag cata-
lyst exhibited the lowest T50 of the different samples tested (Zr5Ag,
Al5Ag). This was attributed to a different distribution of Ag/Ag2O
over the ceria and to the fact that ceria itself is an active catalyst for
soot oxidation.
Reddy et al.159 oxidized soot by a TG method under a ‘tight con-
tact’ condition over CeO2, CeO2-ZrO2 and CeO2-HfO2 nano-oxides.
The observed light-off temperature for the oxidation of model soot
particulates significantly decreased after Zr4+ or Hf4+ ions were incor-
porated into the CeO2 lattice. The improved activity of the doped
catalyst was due to the defective fluorite structure of the ceria with a
concomitant formation of oxygen vacancies.
Cu- and Mn-doped ceria were synthesized by Liang et al.160 by the
sol-gel method. MnOx enters into the CeO2 lattice to form solid solu-
tions, which generate more anion vacancies that adsorb easily releas-
able O2 on the surface. CuxO clusters were highly dispersed on the

b1469_Ch-08.indd 442 4/8/2013 12:35:03 PM


b1469 Catalysis by Ceria and Related Materials

New Developments in Ceria-Based Mixed Oxide Synthesis 443

Figure 8.29. Mechanism of soot combustion with O2 over Fe/CeO2 catalysts.


Reprinted from Zhang et al.157 with permission from Elsevier.

Figure 8.30. Reaction mechanism and active oxygen species. (a) Mn–Ce mixed
oxides under loose contact conditions. (b) Cu–Ce mixed oxides under tight con-
tact conditions. Reprinted from Liang et al.160 with permission from Elsevier.

surfaces of the ceria particles, and this induces a synergetic effect


between these two elements and facilitates more readily releasable
lattice oxygen. A schematic presentation is shown in Fig. 8.30.
Kayama et al.161 evaluated soot oxidation over Ag/CeO2 by TG
analysis of the catalyst in contact with the carbonaceous soot. They

b1469_Ch-08.indd 443 4/8/2013 12:35:03 PM


b1469 Catalysis by Ceria and Related Materials

444 B. M. Reddy, T. Vinod Kumar and N. Durgasri

Figure 8.31 Performance of CeO2−Ag compared with conventional catalysts


and in the absence of a catalyst. (a) Evaluation of carbonaceous soot oxidation.
(b) Arrhenius plots. Reprinted with permission from Kayama et al.161 Copyright
2010 the American Chemical Society.

introduced a novel rice-ball nanocatalyst and observed an interest-


ing activity trend as shown in Fig. 8.31.
Shimizu et al.162 investigated M/CeO2 (M = Ag, Cu, Au and Pt)
catalysts prepared by physical mixing of a metal powder and CeO2
followed by calcination at 500 °C. They pointed out that among the
supported metals, the promotional effect of CeO2 was observed only
for the Ag/CeO2 catalyst. The Ag/CeO2 catalyst showed high activity
for carbon oxidation at lower temperatures as shown in Fig. 8.32, and
was comparable to the same catalyst prepared by impregnation.
Katta et al.163 proposed that promoters with a basic character
show an excellent ability to oxidize soot particles at lower tempera-
tures. Thus a ceria–lanthana (CL) solid solution showed enhanced
soot oxidation over ceria–zirconia despite its lower specific surface
area. The CL sample oxidized 50% of the soot at a lower tempera-
ture than ceria and CZ samples, and the T50 values for ceria, CZ and
CL were 601, 522 and 467 °C, respectively.

8.3.3 CO oxidation
Of the various metal oxides, ceria-based materials have a special sig-
nificance because of their peculiar properties and wide applicability

b1469_Ch-08.indd 444 4/8/2013 12:35:03 PM


b1469 Catalysis by Ceria and Related Materials

New Developments in Ceria-Based Mixed Oxide Synthesis 445

Figure 8.32 Differential thermal analysis (DTA) profiles of carbon oxidation with
20 wt% metal + CeO2 after calcination at 500 °C for 3 h in air. Reprinted from
Shimizu et al.162 with permission from Elsevier.

in various fields of science and technology as mentioned earlier. The


application of ceria-based materials in TWCs is enormous: in 2003
TWC sales accounted for one quarter of the global catalyst market.
Many investigations have accepted that CO oxidation under sta-
tionary conditions occurs over pure ceria by a Mars–van Krevelen
type mechanism, where the reaction involves alternating reduction
and oxidation of the ceria surface with the formation of surface
oxygen vacancies and their successive replenishment by gas-phase
oxygen.164 The limiting steps of CO oxidation catalyzed by ceria via
the Mars–van Krevelen reaction mechanism have been identified
and investigated by means of density functional theory (DFT) calcu-
lations that account for the on-site Coulomb interaction via a
Hubbard term (DFT+U).165
Valechha et al.166 synthesized porous nanostructured ceria by
using chitosan as a biological template. The catalytic activity of vari-
ous ceria-based samples for CO oxidation was evaluated using a
fixed-bed quartz reactor at atmospheric pressure. The T50 for a Pt/3
(1:3 of Ce3+ and chitosan) catalyst was 94 °C and 100% CO conver-
sion was achieved at 160 °C at a space velocity of 60000 h−1. CO
oxidation over various samples prepared under different concentra-
tions (S1 to S7) is shown in Fig. 8.33.

b1469_Ch-08.indd 445 4/8/2013 12:35:04 PM


b1469 Catalysis by Ceria and Related Materials

446 B. M. Reddy, T. Vinod Kumar and N. Durgasri

Figure 8.33 CO oxidation as a function of temperature over different ceria


samples in comparison to commercial ceria. Reprinted from Valechha et al.166 with
permission from the Royal Society of Chemistry.

The CO oxidation activity of ceria (C) catalysts doped by Mg2+


(M), Al3+ (A) and Si4+ (S) followed an interesting trend: CA450 >
CA750 > C450 > CM450 > CS450 > CS750 > CM750 > C750 (the
number represents the calcination temperature). The high catalytic
activity of the CeO2-Al2O3 combination is due to its excellent tex-
tural and structural properties, good homogeneity and redox ability
as shown in Fig. 8.34.167
Reddy et al.168 synthesized ceria-based mixed oxides with various
dopants, namely SiO2, TiO2, ZrO2 and Al2O3, by the microwave-
induced solution combustion method and analysed them for CO
oxidation activity. Of the samples tested, ceria–zirconia exhibited
the highest conversion and lowest light-off temperature (100%,
351 °C), followed by ceria–titania (85%, 404 °C), ceria–alumina
(70%, 454 °C) and ceria–silica (40%), respectively. The incorpora-
tion of zirconia into the ceria lattice facilitates the formation of
more defect sites, which are responsible for the better activity of the
ceria–zirconia mixed oxide.

b1469_Ch-08.indd 446 4/8/2013 12:35:04 PM


b1469 Catalysis by Ceria and Related Materials

New Developments in Ceria-Based Mixed Oxide Synthesis 447

Figure 8.34 CO oxidation activity (at 350 °C) against textural and reducible prop-
erties. Reprinted from Yu et al.167 with permission from Elsevier.

Very recently, Rovira et al.169 synthesized ceria–praseodymia


nanotubes (Ce0.8Pr0.2O2−δ-NT) by employing a template-based elec-
trodeposition method inside the pores of anodic aluminium oxide
(AAO) membranes. CO oxidation at 200 °C over the newly pre-
pared Ce0.8Pr0.2O2−δ-NT/AAO sample was found to be ~1,000 times
faster than over Ce0.8Pr0.2O2−δ obtained by the conventional method,
and 1.5 times higher than over CeO2-NT/AAO. These differences
have been tentatively attributed to the enhancement of redox prop-
erties induced by the incorporation of Pr into the CeO2 lattice and
the influence of the nanostructure. Typical CO oxidation results
over these samples are shown in Fig. 8.35.
Hernández et al.170 synthesized an interesting combination of
Ce1−xEuxO2−x/2 mixed oxides by CP. The europium dopant exhibited
a very strong influence on catalytic activity, and the solid containing
10 wt% Eu2O3 showed the highest activity (100% conversion at
400 °C). Its better activity was again related to a higher number of
oxygen vacancies as confirmed from Raman spectroscopic analysis.
Li et al.171 synthesized Ce1−xTbxO2−δ (CT, x-6.5 at%) nanobelts,
nanoparticles and nanosheets. They were calcined at different tem-
peratures and their OSC properties evaluated under dry air.

b1469_Ch-08.indd 447 4/8/2013 12:35:04 PM


b1469 Catalysis by Ceria and Related Materials

448 B. M. Reddy, T. Vinod Kumar and N. Durgasri

Figure 8.35 CO conversion vs. temperature over various catalysts. Reprinted from
Rovira et al.169 with permission from Elsevier.

Figure 8.36 Conversion of CO over ceria–terbia in the form of (a) nanobelts,


(b) nanosheets and (c) nanoparticles as a function of reaction temperature.
Reprinted from Li et al.171 with permission from the Royal Society of Chemistry.

Interestingly, the nanobelts showed a 96% CO conversion at about


640 °C. However, the nanoparticles and nanosheets showed only
43% and 57% CO conversion at the same temperature (Fig. 8.36).
The high catalytic activity of the nanobelts was attributed to the for-
mation of special nanostructures with high OSC.

b1469_Ch-08.indd 448 4/8/2013 12:35:04 PM


b1469 Catalysis by Ceria and Related Materials

New Developments in Ceria-Based Mixed Oxide Synthesis 449

Figure 8.37 Conversion of CO over CexZr1−xO2 (CZ), CexHf1−xO2 (CH), CexTb1−


xO2−δ (CT) and CexPr1−xO2−δ (CP) solid solutions calcined at 500 °C. Reprinted from
Reddy et al.172 with permission from Springer.

Reddy et al.172 reported that the reducibility of ceria (C) increases


upon doping with Zr4+ (Z), Hf4+ (H), Tb3+/4+ (T) and Pr3+/4+ (P) cati-
ons. The CO oxidation activity was measured for various samples at
normal atmospheric pressure in a fixed-bed microreactor over a
temperature range of 27–500 °C, which revealed that the T50 for the
samples followed the order CH < CT < CP < CZ as shown in Fig. 8.37.
The better activity of the CH sample was correlated with its stable
homogeneous phase formation associated with a high distribution
of oxygen vacancies and higher bulk oxygen mobility with a lower
diffusion energy barrier.
The interest in the use of gold as a catalyst component has been
increasing over the last 15 years, and surprisingly it shows

b1469_Ch-08.indd 449 4/8/2013 12:35:04 PM


b1469 Catalysis by Ceria and Related Materials

450 B. M. Reddy, T. Vinod Kumar and N. Durgasri

Figure 8.38 CO conversion as a function of temperature over different catalysts.


Reprinted with permission from Venezia et al.173 Copyright 2005 the American
Chemical Society.

high activity for CO oxidation at low temperatures. Venezia et al.173


prepared ceria-supported gold catalysts by solvated metal atom dis-
persion (SMAD), DP and CP and evaluated the CO oxidation activ-
ity (Fig. 8.38). Structural and surface analyses evidenced the
presence of a modified ceria phase for the DP sample and the pres-
ence of pure ceria and gold metal crystallites for the SMAD and CP
samples. The presence of metallic gold is expected to weaken the
C–O bond thereby facilitating further insertion of oxygen with
the release of CO2. On the other hand, the presence of ionic gold
in the DP sample increased its activity, which could be attributed to
the weakening of the Ce–O bond and increased oxygen mobility.
Luo et al.174 used surfactant-templated and citrate sol-gel
techniques to prepare CuO/CeO2 samples with a high surface area
and investigated CO oxidation. Their study revealed that finely dis-
persed CuO species over the ceria surface exhibit the highest activity
at 100 °C. The Au–CuOx/CeO2 catalyst was more active in CO

b1469_Ch-08.indd 450 4/8/2013 12:35:04 PM


b1469 Catalysis by Ceria and Related Materials

New Developments in Ceria-Based Mixed Oxide Synthesis 451

conversion than CuOx/CeO2 at low temperatures but exhibited a


similar behaviour at temperatures higher than 120 °C. At 100 °C,
CO conversion decreased in the following order: Au–CuOx/CeO2 >
Au/CeO2 > CuOx/CeO2 with 100%, 98% and 85% conversion,
respectively.175
Panzera et al.176 investigated the effects of calcination tempera-
ture and GHSV over Au/CeOx catalysts for CO conversion. Their
results indicated that calcination at 500 °C results in the establish-
ment of adequate interfacial metal/metal oxide properties, which
are essential for promoting CO oxidation. A significant decrease in
CO conversion was observed with an increase in GHSV, whereas CO2
selectivity slightly increased. The effect of the calcination treatment
on CO conversion over this catalyst is shown in Fig. 8.39.
Rynkowski et al.177 synthesized Au/CZ catalysts using the sol-gel
route followed by direct anionic exchange and evaluated CO oxida-
tion. The activity of an Au/CZ catalyst was found to depend on the
Ce/Zr molar ratio and increased in the following order: Au/
Ce0.25Zr0.75O2 < Au/Ce0.5Zr0.5O2 < Au/Ce0.75Zr0.25O2. Zn-modified ceria

Figure 8.39 Endurance test for uncalcined and calcined Au/CeO2 catalysts.
Reprinted from Panzera et al.176 with permission from Elsevier.

b1469_Ch-08.indd 451 4/8/2013 12:35:04 PM


b1469 Catalysis by Ceria and Related Materials

452 B. M. Reddy, T. Vinod Kumar and N. Durgasri

Figure 8.40 CO oxidation over various gold catalysts. Reprinted from Laguna
et al.178 with permission from Elsevier.

solids were prepared by thermal decomposition and 1 wt% gold was


deposited over the surface by Laguna et al.178 The Zn-modified gold
catalyst was found to be more active for CO oxidation than the
unmodified Au/CeO2, as shown in Fig. 8.40. This behaviour was
proved to be due to a higher amount of metallic dispersed gold on
the Ce–Zn-oxide support surface. The number of oxygen vacancies
acting as nucleation sites for gold were hardly modified in the
Zn-modified ceria support. Therefore, the higher gold dispersion
was related to a high number of electron density sites on the catalyst
surface as a result of Au–Ce–Zn interactions.
Laguna et al.179 synthesized Ce–Fe mixed oxides by a pseudo-sol-
gel method and 1 wt% gold was dispersed by DP. The addition of iron
to the ceria catalyst resulted in an enhancement of the catalytic activity
for CO oxidation especially at low temperatures. The results indicated
that the promoting effect of Au causes 100% CO conversion at tem-
peratures below 100 °C. Another advantage of this catalyst is that it is
stable at 160 °C up to a minimum of 10 h. Various CuO/CZ catalysts
were prepared by wet impregnation with 7 wt% copper and screened
for CO oxidation by Ayastuy et al.180 The CuO/CZ catalyst was able to
oxidize CO completely at 111 °C under a GHSV of 12000 h−1.

b1469_Ch-08.indd 452 4/8/2013 12:35:04 PM


b1469 Catalysis by Ceria and Related Materials

New Developments in Ceria-Based Mixed Oxide Synthesis 453

Alumina-supported CZ catalysts exhibited relatively better per-


formance than titania- and silica-supported samples as reported by
Reddy et al.181 The T50 temperature for CZ/A, CZ/T and CZ/S cata-
lysts was found to be 384, 394 and 497 °C, respectively, indicating the
catalytic efficiency order of CZ/A > CZ/T > CZ/S. In all cases, there
was no appreciable difference in the composition of the Ce–Zr
oxides over the supports except their specific surface area. The
characteristics of various nano-oxide catalysts with their T50 for CO
conversion as reported in the literature are shown in Table 8.3.

8.3.4 Hydrocarbon oxidation


Ceria is one of the better hydrocarbon oxidation catalysts found in
useful products. Recently, Vivier and Duprez186 reviewed various

Table 8.3 CO conversion over different ceria-based materials.


Method of BET surface
Catalyst preparation area (m2.g−1) T50 (°C) Reference
Nano CeO2 Precipitation (ppt) 144.2 250 164
Ce–Tb nanobelts Electrodeposition 115 523 171
Ce0.8Hf0.2O2 CP 78 300 172
Ce0.5Zr0.5O2 Microwave 56 381 168
CuO/Ce0.5Zr0.5O2 Wet impregnation 150 111 180
Ce0.8Pr0.2O2 CP 72 383 21
Ce0.8Pr0.2O2–δ/AAO Electrodeposition — 157 169
Au/CeO2 DP 78.7 225 182
Au/Ce0.8Ga0.2O2 DP 100 111 183
Au/Ce0.9Eu0.1O2 DP 86 128 184
Au/Ce0.9Zn0.1O2 DP 62 50 178
Ce0.75Zr0.25O2/Al2O3 DP 146 384 181
Ce0.75Zr0.25O2/TiO2 DP 105 394 181
Ce0.75Zr0.25O2/SiO2 DP 172 497 181
CeO2-Al2O3 CP 115.9 290 167
CuO-CoO/CeO2-Al2O3 Impregnation 140 91 185
CuO-NiO/CeO2-Al2O3 Impregnation 145 119 185

b1469_Ch-08.indd 453 4/8/2013 12:35:04 PM


b1469 Catalysis by Ceria and Related Materials

454 B. M. Reddy, T. Vinod Kumar and N. Durgasri

organic transformations catalyzed by ceria-based catalysts. Oxygen


reducibility and oxygen storage capacity seem to be important prop-
erties in the performance of ceria in oxidation reactions. The addi-
tion of transition metals, such as Cu and Ag, improved low-temperature
CH4 oxidation activity.187 For the oxidation of ethanol and toluene,
it was observed that catalytic activity was not governed by specific
surface area or metal content but was dependent on the nature of
the metal (gold versus platinum), the nature of the specific support,
the particle oxidation degree and, to a lesser extent, the particle size
of the metal or the support.188
In the 1990s, there were many studies on ceria-based materials,
notably on their synthesis and their catalytic properties. In particu-
lar, the materials were used as structural and electronic promoters
to increase the activity, selectivity and thermal stability of catalysts.
The activity, selectivity and stability of Ni-doped ceria catalysts, with
Ni content ranging from 5 to 20 at%, were investigated for the par-
tial oxidation of CH4 to syngas in the medium-high temperature
range 550–700 °C at atmospheric pressure.186 It was noted that ceria–
titania mixed oxide supported Ni catalysts exhibit high catalytic
activity and stability. Of the various catalysts investigated, the highest
catalytic activity was observed over a 10 wt% NiO/Ce0.5Ti0.5O2 cata-
lyst, which showed excellent stable performance during 100 h on
stream at 750 °C.189 Several studies have revealed that ceria-based
catalysts are also involved in oxidative dehydrogenation reactions190
and in various organic transformations.186

8.4 Conclusions
This chapter has summarized recent progress in the synthetic tech-
niques and application of ceria-based catalysts for combustion and
oxidation reactions. The growing interest in the synthesis of novel
nanosized ceria-based materials is primarily due to their many appli-
cations in different fields. These materials are prepared by both
physical (top-down) and chemical (bottom-up) methods. Compared
to physical methods, the chemical methods play a unique role in
assembling and building ceria-based nanometric units from smaller

b1469_Ch-08.indd 454 4/8/2013 12:35:04 PM


b1469 Catalysis by Ceria and Related Materials

New Developments in Ceria-Based Mixed Oxide Synthesis 455

units. The major goal in the various synthetic methods is to increase


the stability and to enhance the dispersion of nanoparticles in a nar-
row size range by controlling a range of parameters. Recent advances
in the design and production of materials provide tailored sizes,
shapes and distributions, overcoming the main drawbacks of tradi-
tional synthetic methodologies. However, often the synthesis proto-
col for a particular nanomaterial involves not just one, but a
combination of several methods. Ceria-based materials play a very
significant role in combustion processes besides their many practical
applications. In this chapter we discussed VOC abatement and CO,
soot and hydrocarbon oxidation for pollution control on single and
mixed oxides. The catalytic activity and thermal stability of the mate-
rials is very significant in combustion applications. Many studies have
pointed out that the catalytic activity of ceria and modified ceria is
dependent on several factors such as the nature of the support, par-
ticle size, morphology and the interaction between the doped metal
and the support oxide. As of today, the nature of active sites on ceria
modified with other metals, and the reaction mechanisms for CO,
soot, VOCs and hydrocarbons are still the subjects of debate. Much
research has been focussed in recent decades on the catalytic prop-
erties of ceria-based materials, but still there is tremendous scope to
study and exploit these materials for numerous applications.

Acknowledgments
We wish to specially acknowledge all the researchers whose work is
described in this chapter for their valuable contributions. T.V.K. and
N.D.S. are the recipients of a junior research fellowship from CSIR,
New Delhi. Financial support was received from Department of Science
and Technology, New Delhi, under SERC Scheme (SR/S1/PC-63/2008).

References
1. Mirkin, C.A., Small, 1 (2005) 14–16.
2. Campelo, J.M., Luna, D., Luque, R., Marinas, J.M., Romero, A.A.,
Chem Sus Chem., 2 (2009) 18–45.

b1469_Ch-08.indd 455 4/8/2013 12:35:04 PM


b1469 Catalysis by Ceria and Related Materials

456 B. M. Reddy, T. Vinod Kumar and N. Durgasri

3. White, R.J., Luque, R., Budarin, V.L., Clark, J.H., Macquarrie, D.J.,
Chem. Soc. Rev., 38 (2009) 481–494.
4. Goesmann, H., Feldmann, C., Angew. Chem., Int. Ed., 49 (2010)
1362–1395.
5. Nagarajan, R., in Nanoparticles: Synthesis, Stabilization, Passivation, and
Functionalization, ed. R. Nagarajan, American Chemical Society,
Washington, DC, (2008) pp. 1–14, http://users.encs.concordia.ca/
~mojtaba/elec6271/Assignment1-2010.pdf. Accessed 16 June 2011.
6. Bromley, S.T., Moreira, I. de P.R., Neyman, M.K., Illas, F., Chem. Soc.
Rev., 38 (2009) 2657–2670.
7. Ma, Y., Wang, X., Li, S., Toprak, M.S., Zhu, B., Muhammed, M., Adv.
Mater., 22 (2010) 1640–1644.
8. Reddy, B.M., in Metal Oxides: Chemistry and Applications, ed. J.L.G. Fierro,
CRC Press, Florida, (2006) pp. 215–246.
9. Bruix, A., Migani, A., Vayssilov, G.N., Neyman, K.M., Libuda, J., Illas, F.,
Phys. Chem. Chem. Phys., 13 (2011) 11384–11392.
10. Silva, G.A., Nat. Nano Technol., 1 (2006) 92–94.
11. Mallikarjuna, N.N., Thomas, F.S., Rajender, S.V., Acc. Chem. Res., 44
(2011) 469–478.
12. Guo, S., Wang, E., Acc. Chem. Res., 44 (2011) 491–500.
13. Brian, L.C., Vladimir, L.K., O’Connor, C.J., Chem. Rev., 104 (2004)
3893–3946.
14. http://en.wikipedia.org/wiki/Coprecipitation. Accessed 16 June 2011.
15. http://nanospinel.blogspot.in/2007/08/nanopspinel-synthesis-by.
html. Accessed 16 June 2011.
16. Gaikwad, A.B., Navale, S.C., Samuel, V., Murugan, A.V., Ravi, V., Mater.
Res. Bull., 41 (2006) 347–353.
17. Reddy, B.M., Khan, A., Yamada, Y., Kobayashi, T., Loridant, S.,
Volta, J.-C., J. Phys. Chem. B, 107 (2003) 11475–11484.
18. Reddy, B.M., Bharali, P., Saikia, P., Ataullah, K., Stephane, L., Muhler, M.,
Grunert, W., J. Phys. Chem. C, 111 (2007) 1878–1881.
19. Reddy, B.M., Lakshmi, K., Gode, T., Chem. Mater., 22 (2010) 467–475.
20. Reddy, B.M., Saikia, P., Bharali, P., Yusuke, Y., Tetsuhiko, K., Muhler, M.,
Grunert, W., J. Phys. Chem. C, 112 (2008) 16393–16399.
21. Reddy, B.M., Gode, T., Lakshmi, K., Yusuke, Y., Park. S.-E., J. Phys.
Chem. C, 113 (2009) 15882–15890.

b1469_Ch-08.indd 456 4/8/2013 12:35:04 PM


b1469 Catalysis by Ceria and Related Materials

New Developments in Ceria-Based Mixed Oxide Synthesis 457

22. Reddy, B.M., Thrimurthulu, G., Lakshmi, K., Chin. J. Catal., 32 (2011)
800–806.
23. Abimanyu, H., Ahn, B.S., Kim, C.S., Yoo, K.S., Ind. Eng. Chem., Res., 46
(2007) 7936–7941.
24. Wang, Q., Li, Z., Zhao, B., Li, G., Zhou, R., J. Mol. Catal. A: Chem., 344
(2011) 132–137.
25. Feng, S., Pan, D., Wang, Z., Adv. Powder Technol., 22 (2011) 678–681.
26. Reddy, B.M., Saikia, P., Bharali, P., Park. S.-E., Muhler, M., Grunert,
W., J. Phys. Chem. C, 113 (2009) 2452–2462.
27. Shigeyuku, S., Rustum, R., Bull. Mater. Sci., 23 (2000) 453–460.
28. Yoshimura, M., Byrappa, K., J Mater. Sci., 43 (2008) 2085–2103.
29. Patzhe, G.R., Zhou, Y., Kontic, R., Conrad, F., Angew. Chem., Int. Ed.,
50 (2011) 826–859.
30. Yao, W.-T., Yu, S.-H., Int. J. Nanotecchnol., 4 (2007) 129–140.
31. Hayashi, H., Hakuta, Y., Materials, 3 (2010) 3794–3717.
32. Barrer, R.M., Hydrothermal Chemistry of Zeolites, Academic Press,
London, (1982).
33. Feng, S., Xu, R., Acc. Chem. Res., 34 (2001) 239–247.
34. http://scholarbank.nus.edu.sg/bitstream/handle/10635/14809/
03_Chapter%201.pdf?sequence=3. Accessed 16 June 2011.
35. Qin, J., Lu, J., Cao, M., Hu, C., Nanoscale, 2 (2010) 2739–2743.
36. Zhang, Y., Zhang, L., Deng, J., Dai, H., He, H., Inorg. Chem., 48 (2009)
2181–2192.
37. Adschiri, T., Lee, Y.-W., Goto, M., Takami, S., Green Chem., 13 (2011)
1380–1390.
38. Kim, J.-R., Myeong, W.-J., Ihm, S.-K., Appl. Catal. B., 71 (2007) 57–63.
39. Devaraju, M.K., Liu, X., Yusuke, K., Yin, S., Sato, T., Nanotechnol., 20
(2009) 405606.
40. Fernández-González, R., Julián-López, B., Cordoncillo, E., Escribano,
P., J. Mater. Chem., 21 (2011) 497–504.
41. Devaraju, M.K., Shu, Y., Tsugio, S., ACS Appl. Mater. Interfaces, 1 (2009)
2694–2698.
42. http://www.gitam.edu/eresource/nano/nanotechnology/bottamup%
20app.htm. Accessed 16 June 2011.
43. Wen, J., Wilkes, G.L., Chem. Mater., 8 (1996) 1667–1681.
44. Reddy, B.M., Bharali, P., Saikia, P. (2009) In New Nanotechniques, eds.
A. Malik, R.J. Rawat, pp. 243–276.

b1469_Ch-08.indd 457 4/8/2013 12:35:05 PM


b1469 Catalysis by Ceria and Related Materials

458 B. M. Reddy, T. Vinod Kumar and N. Durgasri

45. Liu, J., Cao, G., Yang, Z., Wang, D., Dubois, D., Zhou, X., Graff, G.L.,
Pederson, L.R., Zhang, J., Chem Sus Chem., 1 (2008) 676–697.
46. Livage, J., Henry, M., Sanchez, C., Prog. Solid State Chem., 18 (1988)
259–341.
47. Ward, D.A., Ko, E.I, Ind. Eng. Chem. Res., 34 (1995) 421–433.
48. Carreon, M.A., Guliants, V.V., Eur. J. Inorg. Chem., 1 (2005) 27–43.
49. Inorganic Materials Synthesis, University of Oslo, http://www.uio.no/
studier/emner/matnat/kjemi/KJM5100/h08/undervisningsmateria
le/10KJM5100_20 08_sol_gel_a.pdf. Accessed 16 June 2011.
50. Dimitriev, Y., Ivanova, Y., Iordanova, R., J. Univ. Chem. Technol. Metall.,
43 (2008) 181–192.
51. Bonnetot, B., Rakic, V., Yuzhakova, T., Guimon, C., Auroux, A., Chem.
Mater., 20 (2008) 1585–1596.
52. Tao, Y., Shao, J., Wang, J., Wang, W., J. Alloys Compd., 484 (2009)
729–733.
53. Trinchi, A., Li, Y.X., Wlodarski, W., Kaciulis, S., Pandolfi, L., Viticoli, S.,
Comini, E., Sberveglieri, G., Sens. Actuators, B, 95 (2003) 145–150.
54. Yuan, Q., Liu, Q., Song, W.-G., Feng, W., Pu, W.-L., Sun, L.-D.,
Zhang, Y.-W., Yan, C.-H., J. Am. Chem. Soc., 129 (2007) 6698–6699.
55. Yuan, Q., Duan, H.-H., Li, L.-L., Sun, L.-D., Zhang, Y.-W., Yan, C.-H.,
J. Colloid Interface Sci., 335 (2009) 151–167.
56. Ashok, K.G., Aparna, G., Sonalika, V., Chem., Soc. Rev., 39 (2010)
474–485.
57. Dwars, T., Paetzold, E., Oehme, G., Angew. Chem., Int. Ed., 44 (2005)
7174–7199.
58. Burda, C., Chen, X., Narayanan, R., El-Sayed, M.A., Chem. Rev., 105
(2005) 1025–1102.
59. Capek, I., Adv. Colloid Interface Sci., 110 (2004) 49–74.
60. Adachi, G.-Y., Masui, T., in Catalysis by Ceria Related Materials, ed.
A. Trovarelli, Imperial College Press, London, (2001) pp. 51–76.
61. Lee, J.-S., Lee, J.-S., Choi, S.-C., Mater. Lett., 59 (2005) 395–398.
62. Laguna, O.H., Centeno, M.A., Boutonnet, M., Odriozola, J.A., Appl.
Catal. B: Environ., 106 (2011) 621–629.
63. Bai, J., Xu, Z., Zheng, Y., Yin, H., Mater. Lett., 60 (2006) 1287–1290.
64. http://en.wikipedia.org/wiki/Microwave_oven. Accessed 16 June
2011.

b1469_Ch-08.indd 458 4/8/2013 12:35:05 PM


b1469 Catalysis by Ceria and Related Materials

New Developments in Ceria-Based Mixed Oxide Synthesis 459

65. Gaba, M., Dhingra, N., Ind. J. Pharm. Educ. Res., 45 (2011) 175–183.
66. Galema, S.A., Chem. Soc. Rev., 26 (1997) 233–238.
67. Bilecka, I., Niederberger, M., Nanoscale, 2 (2010) 1358–1374.
68. Liao, X.H., Zhu, J.J., Chen, H.Y., Mater. Sci. Eng. B, 85 (2001) 85–89.
69. Wang, H., Xu, J.-Z., Zhu, J.-J., Chen, H.-Y., J. Cryst. Growth, 244 (2002)
88–94.
70. Godinho, M., Ribeiro, C., Longo, E., Leite, E.R., Crys. Growth Des., 8
(2008) 384–386.
71. Reddy, B.M., Reddy, G.K., Ganesh, I., Ferreira, J.M.F., Catal Lett., 130
(2009) 227–234.
72. Putna, E.S., Vohs, J.M, Gorte, R.J., J. Phys. Chem., 100 (1996)
17862–17865.
73. Bondioli, F., Ferrari, A.M., Leonelli, C., Siligardi, C., Hart, N.A., Evans,
N.G., J. Mater. Chem., 11 (2001) 2620–2624.
74. Bondioli, F., Ferrari, A.M., Lusvarghi, L., Manfredini, T., Nannarone, S.,
Pasqualia, L., Selvaggia, G., J. Mater. Chem., 15 (2005) 1061–1066.
75. Tavakoli, A., Sohrabi, M., Kargari, A., Chem. Pap., 61 (2007) 151–170.
76. Khalil, H., Mahajan, D., Rafailovich, M., Gelfer, M., Pandya, K.,
Langmuir, 20 (2004) 6896–6903.
77. Rodriguez, J.A., Garcia, M.F., eds., Synthesis, Properties, and Applications
of Oxide Nanomaterials, Wiley Interscience, New Jersey, (2007).
78. Bang, J.H., Suslick, K.S., Adv. Mater. 22 (2010) 1039–1059.
79. Zhang, K., Park, B.-J., Fang, F.-F., Choi, H.J., Molecules, 14 (2009)
2095–2110.
80. Shchukin, D.G., Möhwald, H., Phys. Chem. Chem. Phys., 8 (2006)
3496–3506.
81. Yu, J.C., Zhang, L., Lin, J., J. Colloid Interface Sci., 260 (2003)
240–243.
82. Guo, J., Xin, X., Zhang, X., Zhang, S., J. Nanopart. Res., 11 (2009)
737–741.
83. Zhang, D., Fu, H., Shi, L., Pan, C., Li, Q., Chu, Y., Yu, W., Inorg. Chem.,
46 (2007) 2446–2451.
84. Bhattacharyya, S., Gedanken, A., Micropor. Mesopor. Mater., 110 (2008)
553–559.
85. Aruna, S.T., Mukasyan, A.S., Curr. Opin. Solid State Mater. Sci., 12
(2008) 44–50.

b1469_Ch-08.indd 459 4/8/2013 12:35:05 PM


b1469 Catalysis by Ceria and Related Materials

460 B. M. Reddy, T. Vinod Kumar and N. Durgasri

86. Patil, K.C., Aruna, S.T., Mimani, T., Curr. Opin. Solid State Mater. Sci., 6
(2002) 507–512.
87. Chavan, S.V., Tyagi, A.K., J. Mater. Res., 19 (2004) 3181–3188.
88. Tyagi, A.K., Barc. News Lett., (2006) 39–48.
89. Kumar, A., Wolf, E.E., Mukasyan, A.S., AIChE J., 57 (2011) 2207–2214.
90. Mukasyan, A.S., Dinka, P., Int. J. Self Propag. High Temp. Synth., 16
(2007) 23–35.
91. Palmisano, P., Russo, N., Fino, P., Fino, D., Badini, C., Appl. Catal. B:
Environ., 69 (2006) 85–92.
92. Sekar, M.M.A., Manoharan, S.S., Patil, K.C., J. Mater. Sci. Lett., 9 (1990)
1205–1206.
93. Chinarro, E., Jurado, J.R., Colomer, M.T., J. Eur. Ceram. Soc., 27 (2007)
3619–3623.
94. Baidya, T., Marimuthu, A., Hegde, M.S., Ravishankar, N., Madras, G.,
J. Phys. Chem. C, 111 (2007) 830–839.
95. Mukasyan, A.S., Dinka, P., Adv. Eng. Mater., 9 (2007) 653–657.
96. Dinka, P., Mukasyan, A., J. Phys. Chem. B, 109 (2005) 21627–21633.
97. http://www.rsc.org/ebooks/archive/free/CL9780851865546/
CL9780851865546-00001.pdf. Accessed 16 June 2011.
98. Zhao, T., Boullosa-Eiras, S., Yu, Y., Chen, D., Holmen, A., Ronning, M.,
Top. Catal., 54 (2011) 1163–1174.
99. Kakihana, M., Yoshimura, M., Bull. Chem. Soc. Jpn., 72 (1999)
1427–1443.
100. Tang, C.W., Wang, C.B., Chien, S.H., Catal. Lett., 131 (2009) 76–83.
101. Therese, G.H.A., Kamath, P.V., Chem. Mater., 12 (2000) 1195–1204.
102. Karami, H., Mohammadzadeh, E., Int. J. Electrochem. Sci., 5 (2010)
1032–1045.
103. Ou, Y.-N., Li, G.-R., Liang, J.-H., Feng, Z.-P., Tong, Y.-X., J. Phys. Chem.
C, 114 (2010) 13509–13514.
104. Li, G.-R., Qu, D.-L., Tong, Y.-X., J. Phys. Chem. C, 113 (2009) 2704–2709.
105. Wang, Z.-L., Li, G.-R., Ou, Y.-N., Feng, Z.-P., Qu, D.-L., Tong, Y.-X.,
J. Phys. Chem. C, 115 (2011) 351–356.
106. Sen, R., Das, S., Das, K., Mater. Sci. App., 2 (2011) 416–420.
107. Zheng, Y., Hu, Z., Huang, H., Ji, W., Sun, M., Chen, C., J. Nanomater.,
(2011) doi:10.1155/2011/657516.
108. Tsuzuki, T., Mccormick, P.G., J. Mater. Sci., 39 (2004) 5143–5146.

b1469_Ch-08.indd 460 4/8/2013 12:35:05 PM


b1469 Catalysis by Ceria and Related Materials

New Developments in Ceria-Based Mixed Oxide Synthesis 461

109. Fejes, D., Forró, L., Hernadi, K., Phys. Status Solidi B, 247 (2010)
2713–2716.
110. http://www.sigmaaldrich.com/sigma-aldrich/technical-documents/
articles/material-matters/mechanical-processing.html. Accessed 16
June 2011.
111. De Castro, C.L., Mitchell, B.S., in Synthesis, Functionalization and
Surface Treatment of Nanoparticles, ed. M.-I. Baraton, American Scientific
Publishers, Valencia, CA, (2002), pp. 1–15.
112. Lia, Y.X., Zhou, X.Z., Wang, Y., You, X.Z., Mater. Lett., 58 (2003) 245–249.
113. Hos, J.P., McCormick, P.G., Scr. Mater., 48 (2003) 85–90.
114. Trovarelli, A., Zamar, F., Llorca, J., de Leitenburg, C., Dolcetti, G.,
Kiss, J.T., J. Catal., 169 (1997) 490–502.
115. Zhang, T.S., Ma, J., Kong, L.B., Hing, P., Leng, Y.J., Chan, S.H.,
Kilner, J.A., J. Power Sour., 124 (2003) 26–33.
116. Rupp, J.L.M., Infortuna, A., Gauckler, L.J., J. Am. Ceram. Soc., 90
(2007) 1792–1797.
117. Matthey, J., Platinum Met. Rev., 55 (2011) 149–151.
118. Jung, D.S., Park, S.B., Kang, Y.C., Korean J. Chem. Eng., 27 (2010)
1621–1645.
119. Chen, C.Y., Tseng, T.K., Tsai, S.C., Lin, C.K., Lin, H.M., Ceram. Inter.,
34 (2006) 409–416.
120. Min Kim., PhD Thesis, University of Michigan (2008).
121. Daniel, B., Danick, B., André, R.S., Nicolaas, F.D., Ludwig, J.G., Adv.
Mater., 18 (2006) 3015–3018.
122. Pichestapong, P., Injarean, U., J. Met. Mater. Miner., 20 (2010) 51–54.
123. Rupp, J.L.M., Drobek, T., Rossi, A., Gauckler, L.J., Chem. Mater., 19
(2007) 1134–1142.
124. Chaisuk, C., Wehatoranawee, A., Preampiyawat, S., Netiphat, S.,
Shotipruk, A., Panpranot, J., Jongsomjit, B., Mekasuwandumrong, O.,
Ceram. Int., 37 (2011) 1459–1463.
125. Fino, D., Russo, N., Saracco, G., Specchia, V., J. Catal., 217 (2003)
367–375.
126. Choudhary, T.V., Banerjee, S., Choudhary, V.R., Appl. Catal. A., 234
(2002) 1–23.
127. Gallardo, S.M., PhD Thesis, Lappeenranta University of Technology
(2008), http://www.umad.de/infos/cleanair13/pdf/full_142.pdf.
Accessed 16 June 2011.

b1469_Ch-08.indd 461 4/8/2013 12:35:05 PM


b1469 Catalysis by Ceria and Related Materials

462 B. M. Reddy, T. Vinod Kumar and N. Durgasri

128. Li, W.B., Wang, J.X., Gong, H., Catal. Today, 148 (2009) 81–87.
129. Enga, B.B.E., Thompson, D.T., Platinum Met. Rev., 23 (1979)
134–141.
130. Thevenin, P., PhD Thesis, Royal Institute of Technology, Stockholm,
(2002).
131. http://www.netl.doe.gov/technologies/coalpower/turbines/ref-
shelf/handbook/3.2.2.pdf. Accessed 16 June 2011.
132. Saint-Just, J., Kinderen, J., Catal. Today, 29 (1996) 387–395.
133. Adamopoulos, O., PhD Thesis, Royal Institute of Technology, Stockholm,
(2003).
134. Rivas, B., Gutiérrez-Ortiz, J.I., López-Fonseca, R., González-Velasco,
J.R., Appl. Catal. A: Gen., 314 (2006) 54–63.
135. Yang, Y., Xu, X., Sun, K., Catal. Commun., 7 (2006) 756–760.
136. Liotta, L.F., Carlo, G.D., Pantaleo, G., Deganello, G., Catal. Commun.,
6 (2005) 329–336.
137. Bozo, C., Guilhaume, N., Garbowski, E., Primet, M., Catal. Today, 59
(2000) 33–45.
138. Wang, J., Liu, Z., Cao, H., Gong, M., Chen, Y., Chen, Y., Chem. Res.
Chin. Univ., 25 (2009) 81–85.
139. Alifanti, M., Florea, M., Pârvulescu, V.I., Appl. Catal. B: Environ., 70
(2007) 400–405.
140. Hu, C., Zhu, Q., Jiang, Z., Chen, L., Wu, R., Chem. Eng. J., 152 (2009)
583–590.
141. Hu, C., Chem. Eng. J., 168 (2011) 1185–1192.
142. Ontelli, G.C., Reolon, R.P., Alves, A.K., Berutti, F.A., Bergmann, C.P.,
Appl. Catal. A: Gen., 405 (2011) 79–83.
143. Scirè, S., Minicò, S., Crisafulli, C., Satriano, C., Pistone, A., Appl. Catal.
B: Environ., 40 (2003) 43–49.
144. Li, J., Li, W., J. Rare Earths, 28 (2010) 547–551.
145. Centeno, M.A., Paulis, M., Montes, M., Odriozola, J.A., Appl. Catal. A:
Gen., 234 (2002) 65–78.
146. Scirè, S., Riccobene, P.M., Crisafulli, C., Appl. Catal. B: Environ., 101
(2010) 109–117.
147. Wang, X., Kang, Q., Li, D., Catal. Commun., 9 (2008) 2158–2162.
148. Wu, M., Wang, X., Dai, Q., Gu, Y., Li, D., Catal. Today, 158 (2010)
336–342.
149. Zarur, A.J., Ying, J.Y., Nature, 403 (2000) 65–67.

b1469_Ch-08.indd 462 4/8/2013 12:35:05 PM


b1469 Catalysis by Ceria and Related Materials

New Developments in Ceria-Based Mixed Oxide Synthesis 463

150. Pez, R.R., Balderas-Tapia, L., Elizalde-Martinez, I., Viveros, T., Chem.
Eng. Commun., 196 (2009) 1189–1197.
151. Aneggi, E., Leitenburg, C., Dolcetti, G., Trovarelli, A., Catal. Today,
114 (2006) 40–47.
152. Hensgen, L., Stöwe, K., Catal. Today, 159 (2011) 100–107.
153. Fang, P., Luo, M.-F., Lu, J.-Q., Cen, S.-Q., Yan, X.Y., Wang, X.-X.,
Thermochim. Acta, 478 (2008) 45–50.
154. Simonsen, S.B., Dahi, S., Johnson, E., Molenbroek, A.M., Helveg, S.,
J. Catal., 255 (2008) 1–5.
155. Aneggi, E., Leitenburg, C., Trovarelli, A., Catal. Today, 181 (2012)
108–115.
156. Atribak, I., López-Suárez, F., Bueno-López, A., García-García, A.,
Catal. Today, 176 (2011) 404–408.
157. Zhang, Z., Han, D., Wei, S., Zhang, Y., J. Catal., 276 (2010) 16–23.
158. Aneggi, E., Llorca, J., Leitenburg, C., Dolcetti, G., Trovarelli, A., Appl.
Catal. B: Environ., 91 (2009) 489–498.
159. Reddy, B.M., Bharali, P., Thrimurthulu, G., Saikia, P., Katta, L.,
Park, S.-E., Catal. Lett., 123 (2008) 327–333.
160. Liang, Q., Wu, X., Weng, D., Xu, H., Catal. Today, 139 (2008) 113–118.
161. Kayama, T., Yamazaki, K., Shinjoh, H., J. Am. Chem. Soc., 132 (2010)
13154–13155.
162. Shimizu, K.I., Kawachi, H., Komai, S.-I. Yoshida, K., Sasaki, Y.,
Satsuma, A., Catal. Today, 175 (2011) 93–99.
163. Katta, L., Sudarsanam, P., Thrimurthulu, G., Reddy, B.M., Appl. Catal.
B: Environ., 101 (2010) 101–108.
164. Aneggi, E., Llorca, J., Boaro, M., Trovarelli, A., J. Catal., 234 (2005) 88–95.
165. Huang, M., Fabris, S., J. Phys. Chem. C, 112 (2008) 8643–8648.
166. Valechha, D., Lokhande, S., Klementova, M., Subrt, J., Rayalu, S.,
Labhsetwar, N., J. Mater. Chem., 21 (2011) 3718–3725.
167. Yu, Q., Wu, X., Tang, C., Qi, L., Liu, B., Gao, F., Sun, K., Dong, L.,
Chen, Y., J. Colloid and Interface Sci., 354 (2011) 341–352.
168. Reddy, B.M., Reddy, G.K., Ganesh, I., Ferreira, J.M.F., J. Mater. Sci., 44
(2009) 2743–2751.
169. Rovira, L.G., Delgado, J.J., ElAmrani, K., Rio, E., Chen, X., Calvino, J.J.,
Botana, F.J., Catal. Today, 180 (2012) 167–173.
170. Hernández, W.Y., Centeno, M.A., Romero-Sarria, F., Odriozola, J.A.,
J. Phys. Chem. C, 113 (2009) 5629–5635.

b1469_Ch-08.indd 463 4/8/2013 12:35:05 PM


b1469 Catalysis by Ceria and Related Materials

464 B. M. Reddy, T. Vinod Kumar and N. Durgasri

171. Li, G.-R., Qu, D.-L., Wang, Z.-L., Su, C.-Y., Tong, Y.-X., Arurault, L.,
Chem. Commun., 48 (2009) 7557–7559.
172. Reddy, B.M., Thrimurthulu, G., Katta, L., Catal. Lett., 141 (2011)
572–581.
173. Venezia, A.M., Pantaleo, G., Longo, A., Carlo, G.D., Casaletto, M.P.,
Liotta, F.L., Deganello, G., J. Phys. Chem. B, 109 (2005) 2821–2827.
174. Luo, M.-F., Song, Y.-P., Lu, J.-Q., Wang, X.-Y., Pu, Z.-Y., J. Phys. Chem. C,
111 (2007)12686–12692.
175. Fonseca, J.S.L., Ferreira, H.S., Bion, N., Pirault-Roy, L., Rangel, M.C.,
Duprez, D., Epron, F., Catal. Today, 180 (2012) 41–43.
176. Panzera, G., Modafferi, V., Candamano, S., Donato, A., Frusteri, F.,
Antonucci, P.L., J. Power Sources, 135 (2004) 177–183.
177. Rynkowski, J.M., Dobrosz-Gómez, I., Annales UMCS, 64 (2009) 197–217.
178. Laguna, O.H., Centeno, M.A., Romero-Sarria, F., Odriozola, J.A.,
Catal. Today, 172 (2011) 118–123.
179. Laguna, O.H., Centena, M.A., Arzamendi, G., Gandía, L.M., Romero-
Sarria, F., Odriozol, J.A., Catal. Today, 157 (2010) 155–159.
180. Ayastuy, J.L., Gurbani, A., González-Marcos, M.P., Gutiérrez-Ortiz, M.A.,
Appl. Catal. A: Gen., 387 (2010) 119–128.
181. Reddy, B.M., Lakshmanan, P., Bharali, P., Saikia, P. Thrimurthulu, G.,
Muhler, M., Grunert, W., J. Phys. Chem. C, 111 (2007) 10478–10483.
182. Lai, S.-Y., Qiu, Y., Wang, S., J. Catal., 237 (2006) 303–313.
183. Vecchietti, J., Collins, S., Delgado, J.J., Małecka, M., Rio, E., Chen, X.,
Bernal, S., Bonivardi, S., Top Catal., 54 (2011) 201–209.
184. Hernández, W.Y., Romero-Sarria, F., Centeno, M.A., Jose Odriozola, A.,
J. Phys. Chem. C, 114 (2010) 10857–10865.
185. Reddy, B.M., Rao, K.N., Bharali, P., Ind. Eng. Chem. Res., 48 (2009)
8478–8486.
186. Vivier, L., Duprez, D., Chem Sus Chem., 3 (2010) 654–678.
187. Kundakovic, L., Flytzani-Stephanopoulos, M., J. Catal., 179 (1998)
203–222.
188. Gaálová, J., Topka, P., Kaluža, L., Šolcová, O., Catal. Today, 175 (2011)
231–237.
189. Zhang, Y., Li, Z., Wen, X., Liu, Y., Chem. Eng. J., 121 (2006) 115–123.
190. Rao, K.N., Reddy, B.M., Park, S.-E., Appl. Catal. B: Environ., 100 (2010)
472–479.

b1469_Ch-08.indd 464 4/8/2013 12:35:05 PM


b1469 Catalysis by Ceria and Related Materials

CHAPTER 9

DESIGN AND MODELING OF ACTIVE


SITES IN METAL–CERIA CATALYSTS
FOR THE WATER GAS SHIFT REACTION
AND RELATED CHEMICAL PROCESSES
Jose A. Rodriguez
Chemistry Department, Brookhaven National Laboratory,
Upton, NY 11973, USA

9.1 Introduction: Ceria-Based Catalysts


for the Water Gas Shift Reaction
This chapter reviews a series of studies examining the water gas shift
reaction on model ceria-based catalysts. Currently, the primary
source of hydrogen for the chemical and petrochemical industries
comes from the steam reforming of hydrocarbons: CnHm + nH2O →
nCO + (n−m/2)H2.1 The reformed fuel usually contains 1–10% of
CO, an impurity that can be a serious problem for chemical pro-
cesses that use H2 as a feedstock and for the operation of fuel cells.
CO oxidation (CO + 0.5O2 → CO2) and the water gas shift (WGS)
reaction (CO + H2O → H2 + CO2) are critical processes for provid-
ing clean hydrogen.1,2 Common industrial catalysts for the WGS
(mixtures of Fe–Cr or Zn–Al–Cu oxides) are pyrophoric and nor-
mally require lengthy and complex activation steps before usage.1
Recent works report that catalysts combining CeO2 with metals
such as Au, Cu, and Pt are very efficient catalysts for the WGS
reaction.2–5

465

b1469_Ch-09.indd 465 4/8/2013 12:36:22 PM


b1469 Catalysis by Ceria and Related Materials

466 J. A. Rodriguez

The nature of the active phase(s) in these metal/oxide WGS


catalysts and the WGS reaction mechanism are still unclear.5 For
example, the as-prepared Au–CeO2 catalysts contain nanoparticles
of pure gold and gold oxides (AuOx) dispersed on a nanoceria
support.3 Each of these gold species could be in the active phase3,5
and the ceria support may not be a simple spectator in these
systems.6,7 Although pure ceria is a very poor WGS catalyst, the
properties of this oxide were found to be crucial for the observed
activity of Au–CeO2 nanocatalysts.2,3,8 Several studies dealing with
metal/oxide powder catalysts and the WGS indicate that the oxide
plays a direct role in the reaction,2,3,5,9–11 but because of the com-
plex nature of these systems, there is no agreement on its role.
Although Au–CeO2 catalysts have a high WGS activity, neither gold
nor ceria alone are able to catalyze the reaction. The results of
density functional (DF) calculations point to a very high barrier for
the dissociation of H2O on Au(111) or Au(100),12 which leads to
negligible activity for the WGS process. Even gold nanoparticles
cannot dissociate water and catalyze the WGS.12 Furthermore, sur-
faces and nanoparticles of copper are by far much better catalysts
for the WGS than surfaces and nanoparticles of gold.12 Thus, how
can Au/CeO2 catalysts be much more active than conventional
Cu/ZnO catalysts?
In this chapter, we will examine recent published studies investi-
gating the behavior of ceria-based WGS catalysts. First, we will focus
on the in situ characterization of gold–ceria, copper–ceria, and
platinum–ceria powder catalysts using time-resolved X-ray diffrac-
tion (XRD) and X-ray absorption spectroscopy.5,10,13 This will be fol-
lowed by a review of studies that investigate the properties of model
WGS catalysts generated by the deposition of Cu and Au nanoparti-
cles on a well-defined CeO2(111) surface.4 Then, we will present
mechanistic studies for the WGS on inverse CeOx/Au(111) and
CeOx/Cu(111) model catalysts.14–17 The chapter ends with a discus-
sion of the properties of novel and extremely active M/CeOx/
TiO2(110) (M=Cu, Au, or Pt) catalysts, which take advantage of the
special properties of ceria nanoparticles supported on a titania
support.18,19

b1469_Ch-09.indd 466 4/8/2013 12:36:23 PM


b1469 Catalysis by Ceria and Related Materials

Design and Modeling of Active Sites in Metal–Ceria Catalysts 467

9.2 Active Phase of Au–CeO2, Cu–CeO2, and Pt–CeO2


Powder Catalysts
In Fig. 9.1, the results of in situ time-resolved XRD show diffraction
lines for CuO and CeO2 in a freshly prepared copper–ceria nano-
catalyst.20 Once the catalyst is exposed to the reactants of the WGS at
temperatures above 200 °C, the diffraction lines for CuO disappear
and new lines appear for metallic Cu. Significant catalytic activity is

CeO2
Cu

500oC

CuO 400oC

300oC
/h
Time

o
200 C


(a)
o
400 C o
500 C
5%Cu/CeO 2
H2 relative concentraion

o
300 C
o
400 C
o Cu0.2Ce 0.8O 2
300 C

o
200 C
o
o
500 C
400 C
300 C
o
CuO

0 2 4 6 8 10 12 14
Time / hour
(b)

Figure 9.1 (a) In situ time-resolved XRD patterns collected for a 5% CuO/CeO2
catalyst during the WGS reaction.20 (b) Amount of H2 produced at several tempera-
tures for three different catalysts containing Cu.20 Copyright 2006 the American
Chemical Society.

b1469_Ch-09.indd 467 4/8/2013 12:36:23 PM


b1469 Catalysis by Ceria and Related Materials

468 J. A. Rodriguez

only observed at temperatures higher than 200 °C when Cu and


partially reduced CeO2 are present.20 The nanocatalyst used for the
experiments in Fig. 9.1 was prepared by impregnating CeO2 with
CuO.20 The two oxides can form solid solutions. The Ce1−xCuxO2
nanocatalysts were prepared by a reverse microemulsion method.21
The lattice of the Ce1−xCuxO2 systems still adopts a fluorite-type
structure like pure ceria, and it is highly stable21 but, nevertheless,
the copper is reduced and comes out during the WGS reaction.13,20
Data of X-ray absorption near-edge spectroscopy (XANES)
indicate that AuOx and PtOx also do not survive under WGS reaction
conditions.5,22 Originally, it was suggested that the active phase in Au–
CeO2 catalysts are AuOx nanoparticles or more specifically cationic
Auδ+ species.3 Au L3-edge XANES spectra collected at room tempera-
ture for fresh Au–CeO2 catalysts show a clear feature at ∼ 2.5 eV above
the edge, which is not seen for metallic gold and is characteristic of
gold oxides.22 The intensity of this peak is higher than that observed
for Au2O and closer to that seen in Au2O3.22 Once the Au–CeO2 cata-
lyst was exposed to a mixture of CO and H2O at elevated tempera-
tures, the XANES features for gold oxide disappeared. At temperatures
above 200 °C, when significant WGS activity was detected, the line
shape of the Au L3-edge resembled that of pure gold.22 Similar results
were found when the ratio of CO/H2O in the feed was changed or
when the ceria support was doped with Zr.5,22 Thus, the in situ time-
resolved XANES data indicate that cationic Auδ+ species cannot be the
key sites responsible for the WGS activity, because they do not exist
under reaction conditions.5,22 An identical finding has been reported
for Pt–CeO2 powder catalysts.5,19 In all these catalysts, the active phase
consisted of small metal aggregates (< 2 nm in size) dispersed on
partially reduced ceria.5,19,22 Figure 9.2 displays Ce L3-edge XANES
spectra for the WGS on a Ce0.8Cu0.2O2 catalyst.13,20 The Ce L3-edge data
indicate that a fraction of the Ce4+ cations in ceria undergoes partial
reduction to Ce3+ during the WGS. For example, the line shape of the
spectrum at 300 °C is different from that expected for nearly stoichio-
metric CeO2 (line shape at 25 °C). The concentration of Ce3+ defects
increases when going from 150 to 400 °C. This is important because
O vacancies in ceria can facilitate the dissociation of H2O.13,20

b1469_Ch-09.indd 468 4/8/2013 12:36:23 PM


b1469 Catalysis by Ceria and Related Materials

Design and Modeling of Active Sites in Metal–Ceria Catalysts 469

Figure 9.2 Ce L3-edge XANES spectra collected for Ce0.8Cu0.2O2 during the WGS
reaction at different temperatures.13,20 The spectrum for Ce2O3 (Ce3+) is also
included for comparison. Copyright 2006 the American Chemical Society.

After establishing the active phase in Cu–CeO2, Au–CeO2, and


Pt–CeO2 powder catalysts,5,19,20 research has been performed using
well-defined model catalysts to investigate fundamental aspects of
the WGS reaction: roles of the metal and oxide in the process,
effects of particle size, and the reaction mechanism. As shown in
Fig. 9.3, model systems with different configurations have been used
to address different issues associated with the WGS. In a conven-
tional model catalyst, nanoparticles of a metal are deposited on top
of a well-defined oxide surface.23,24 In general, this is a reasonable
representation of the configuration seen in most metal/oxide pow-
der catalysts. Following this approach, one can study variations in
catalytic activity as a function of metal particle size, and identify
reaction intermediates.23 On an inverse oxide/metal catalyst, the
reactants can interact with defect sites of the oxide nanoparticles,
metal sites, and the metal/oxide interface.14,15,17
Defect sites present in the oxide are not covered by metal parti-
cles, as in a conventional metal/oxide catalyst. The oxide/metal
system is an attractive model to investigate the role of the oxide in
a catalytic process, and also can be used to study reaction mechanisms

b1469_Ch-09.indd 469 4/8/2013 12:36:23 PM


b1469 Catalysis by Ceria and Related Materials

470 J. A. Rodriguez

Figure 9.3 Different configurations of model catalysts used for studying the WGS.

in detail.14–17 The most complex and promising configuration in


Fig. 9.3 is the mixed metal oxide configuration, in which nanopar-
ticles of a metal and an oxide can interact with the reactants. We will
deal with this kind of catalyst in the final part of this chapter.

9.3 Water Gas Shift Reaction on Au and Cu Nanoparticles


Supported on a Well-defined CeO2(111) Surface
The kinetics of the WGS reaction have been investigated in detail on
model catalysts generated by vapor-depositing nanoparticles of gold
or copper on CeO2(111).4 On this oxide substrate, the admetals grow
and form three-dimensional (3D) particles.4,25 Figure 9.4 displays the
behavior of the Au/CeO2(111) and Cu/CeO2(111) catalysts as a func-
tion of admetal coverage.4 For comparison, we also include the cor-
responding data for the deposition of Au and Cu on ZnO(000ī).4
CeO2(111) and ZnO(000ī) are both O-terminated surfaces. These
oxide supports are inactive as catalysts for the WGS reaction. In Fig.
9.4(a), it can be seen that the catalytic activity of the metal/oxide sys-
tems initially increases when Au is added, reaching a maximum at ∼
0.4–0.5 monolayer (ML). Above these coverages, overall catalytic activ-
ity decreases. Similar trends have been observed for the WGS on Au/
TiO2(110)26 and Au/MoO2.27 Scanning tunneling microscopy (STM)
images show that the particle size of Au on CeO2(111), ZnO(000ī), or
TiO2(110) rises above 4 nm and continuously grows when the admetal
coverage is increased beyond 0.5 ML.4,18,25 The trends in Fig. 9.4(a)

b1469_Ch-09.indd 470 4/8/2013 12:36:23 PM


b1469 Catalysis by Ceria and Related Materials

Design and Modeling of Active Sites in Metal–Ceria Catalysts 471

Figure 9.4 WGS activity of model catalysts generated by vapor-depositing Au (a)


and Cu (b) on CeO2(111) and ZnO(000ī).4 The catalytic activity is reported as a
function of admetal coverage. Each surface was exposed to a mixture of 20 torr of
CO and 10 torr of H2O at 625 K for 5 min. The steady state was reached 2–3 minutes
after introducing the gases into the batch reactor.4 Copyright 2007 Wiley.

b1469_Ch-09.indd 471 4/8/2013 12:36:23 PM


b1469 Catalysis by Ceria and Related Materials

472 J. A. Rodriguez

probably reflect changes in the size of the gold particles: high catalytic
activity is seen for small gold particles (size < 4 nm), and it decreases
as the particle size increases. Although the optimum WGS activity in
Fig. 9.4 is for admetal coverages of 0.4–0.5 ML, metal/oxide catalysts
with gold coverages near 1 ML are still substantially more active than
Au(111) or polycrystalline gold, surfaces that are not catalytically
active.4 For the deposition of Cu on CeO2(111) and ZnO(000ī), one
sees trends that are similar to those found in the case of Au deposi-
tion, but the admetal coverage for higher WGS is somewhat different.
The most active Au/CeO2(111) and Cu/CeO2(111) catalysts have
comparable activity.28 This is quite remarkable because extended sur-
faces of pure copper and pure gold exhibit quite different activities
for the WGS,4,12 with bulk metallic gold being inactive. Thus, the ceria
probably plays an active role behind the excellent performance of
Au/CeO2 catalysts.4
The gold and copper atoms on the CeO2(111) and ZnO(000ī)
surfaces were probably not oxidized during the WGS process.4 After
reaction, X-ray photoelectron spectroscopy (XPS) showed Au 4f and
Cu 2p positions that were almost identical to those seen upon depo-
sition of gold and Cu on the oxides and were very different from
those typically seen for AuOx and CuOx species.4,10 Post-reaction
surface analysis also showed the presence of formate-like and car-
bonate-like groups on the surface of the catalysts. Possible reaction
paths for the formation of these groups are discussed in Burch.5 It is
not completely clear if they are key intermediates in the WGS pro-
cess or simple spectators.4,5,27
Figure 9.5(a) compares the WGS activity of Au(111) and 0.5 ML
of Au supported on ZnO(000ī),4 CeO2(111),4 and an MoO2 film.27
The nature of the support plays a key role in the activation of the
gold nanoparticles. Zinc oxide is frequently used in industrial Cu–
ZnO WGS catalysts.1,5 However, the Au/ZnO(000ī) system displays
low WGS activity when compared to Au/CeO2(111). A similar trend
is found when comparing the activities of Cu/CeO2(111) and Cu/
Zn O(000ī), see Fig. 9.4. The ceria contains a substantial number of
metal cations not fully oxidized under WGS reaction conditions and
may participate directly in important steps of the process.4 This is
not the case for ZnO.4

b1469_Ch-09.indd 472 4/8/2013 12:36:24 PM


b1469 Catalysis by Ceria and Related Materials

Design and Modeling of Active Sites in Metal–Ceria Catalysts 473

Figure 9.5 (a) WGS activity of Au(111) and 0.5 ML of Au supported on


ZnO(000ī),4 CeO2(111),4 and polycrystalline MoO227 (20 torr of CO and 10 torr of
H2O at 625 K for 5 min). (b) Corresponding apparent activation energies.

b1469_Ch-09.indd 473 4/8/2013 12:36:24 PM


b1469 Catalysis by Ceria and Related Materials

474 J. A. Rodriguez

Figure 9.5(b) displays the apparent activation energies observed


for 0.5 ML of Au deposited on ZnO(000ī),4 CeO2(111),4 and
MoO2.27 All the reported values were obtained from Arrhenius
plots, which contained reaction rates measured at 575, 600, 625,
and 650 K under steady-state conditions at PCO = 20 torr and PH2O =
10 torr. The more active catalysts have apparent activation energies
in the range of 7–8.5 kcal.mol−1, while the corresponding value for
the less active catalysts is ∼ 16 kcal.mol−1. Cu deposited on CeO2(111)
or MoO2 also has apparent activation energies smaller than that
seen on ZnO(000ī).4,27 The presence of O vacancies in the
CeO2(111) and MoO2 substrates facilitates the dissociation of water
and, thus, probably leads to a relatively low apparent activation
energy for the WGS.4,27 The next section analyzes the causes behind
the strong effects that ceria has on the WGS activity of gold and
copper.

9.4 Inverse CeOx/Au(111) and CeOx/Cu(111) Catalysts


and the Chemistry Associated with the Water Gas
Shift Reaction
There is a general desire to improve the configuration of industrial
catalysts to take advantage of the intrinsic properties of metal
oxides. In recent years, a series of studies has been published exam-
ining the growth of oxide nanoparticles on metal substrates.9,16,29–42
These studies have revealed structures for the supported oxide,
which are different from those found in bulk phases. In addition,
the oxide↔metal interactions can alter the electronic states of the
oxide producing new chemical properties. On an inverse oxide/
metal catalyst, the reactants can interact with defect sites of the
oxide nanoparticles, metal sites, and the metal/oxide interface. In
these systems, one can couple the special reactivity of the oxide
nanoparticles to the reactivity of the metal to obtain high catalytic
activity. Furthermore, an oxide/metal system is also an attractive
model for fundamental studies. It can be used to investigate the role
of the oxide in a catalytic process, and how the stability of different
reaction intermediates depends on the nature of the oxide.16

b1469_Ch-09.indd 474 4/8/2013 12:36:24 PM


b1469 Catalysis by Ceria and Related Materials

Design and Modeling of Active Sites in Metal–Ceria Catalysts 475

The deposition of ceria nanoparticles and films has been stud-


ied on close-packed metal surfaces of Ni,29 Cu,30,31 Rh,32 Ru,33,34 Pd,38
Re,36 Pt,37–41 and Au.9,42 Well-ordered films were characterized at the
atomic scale with STM for Cu(111),17,31 Ru(001),33,34 Rh(111),32
Pt(111),38–41 and Au(111).9,42 Substantial differences between the lat-
tice constants of the close-packed surfaces of ceria and the metals
listed above seem to play a minor role in the growth of the ordered
oxide overlayer. At the submonolayer level, the morphology, compo-
sition, and reactivity of the ceria nanoparticles depends on the
experimental conditions used for their preparation: oxidation of Ce
overlayers or surface alloys, codeposition of Ce and O2, and reaction
of Ce with multilayers of H2O or NO2 supported on a metal.9,16,29–42
Figure 9.6 shows STM images obtained after depositing ceria
nanoparticles on Au(111) using two different procedures.9,42 In the
first procedure, Fig. 9.6(a), the oxide nanoparticles were synthe-
sized by exposing a Ce/Au(111) alloy surface to O2 gas at 450 K.
There is a random distribution of the ceria nanoparticles on the
terraces of the gold substrate, which maintains its characteristic
herringbone reconstruction.42 The particles occupy specific sites
with respect to the dislocation ridges of the reconstruction.9,42 One
can see large elongated particles with an aspect ratio (mean
length/mean width) > 3. The surface of these ceria nanoparticles is
quite rough. In the second procedure, Fig. 9.6(b), cerium atoms
were vapor-deposited under an atmosphere of O2 at 550 K. This led
to an exclusive decoration of steps (see arrows in Fig. 9.6(b)) by
individual crystalline islands of distorted hexagonal shape, which
are flat and about 30 nm across.42 The islands seem to grow from
the lower step edges; although they are locally embedded in steps
their interaction with the gold is not strong since the herringbone
pattern on the terraces is preserved. The first layer may be a double
layer [O–Ce–O…O–Ce–O/Au] of CeO2 with a typical 0.31 nm dis-
tance between equivalent layers of (111) planes. However, addi-
tional layers grow in increments of 0.23 nm instead of 0.31 nm and
this discrepancy may reflect the electronic structure differences
between the metallic gold step height used as a topographical refer-
ence and the height of the oxide islands.42 Atomically resolved

b1469_Ch-09.indd 475 4/8/2013 12:36:24 PM


b1469 Catalysis by Ceria and Related Materials

476 J. A. Rodriguez

Figure 9.6 (a) STM image obtained after annealing a Ce–Au(111) surface alloy at
450 K in 1 × 10−8 torr of O2 for 30 min and flashed to 650 K. (b) STM image
recorded after depositing Ce on Au(111) at 550 K in 1.5 × 10−7 torr of O2. Taken
from Ma et al.42 Both graphs are differential images. Copyright 2007 Elsevier.

images showed an oxygen-terminated surface of CeO2(111) with


oxygen vacancies.9,42
STM images indicate that ceria also grows on Rh(111) forming
ordered CeO2(111) islands with their (111) faces parallel and orien-
tationally aligned to the main azimuthal directions of the substrate.43
Strain caused by the mismatch between the lattices of the oxide and
metal facilitated the formation of O vacancies in ceria.32 A similar

b1469_Ch-09.indd 476 4/8/2013 12:36:24 PM


b1469 Catalysis by Ceria and Related Materials

Design and Modeling of Active Sites in Metal–Ceria Catalysts 477

phenomenon probably occurs in the CeO2/Au(111) and CeO2/


Cu(111) systems.17,42 After dosing Ce to Cu(111) under an atmos-
phere of O2,17 the copper is oxidized to Cu2O and, on top of this
oxide, one sees two types of ceria features (Fig. 9.7). Small islands of
ceria (2–5 nm in size, CeOx-I) appear on the terraces of the copper
substrate, whereas large islands of ceria (30–50 nm in size, triangu-
lar shape, CeOx-II) are embedded in the substrate step edges. Inside
the large ceria islands, one can see a moiré pattern with a separation
of ∼ 5 nm in between the depressions.17 The large ceria islands have
a structure different from that seen for the two most stable surfaces
of bulk ceria: CeO2(111) and CeO2(110). The oxide nanoparticles
in ceria/Cu(111) display a morphology quite different from that
found for the growth of ceria nanoparticles on an Au(111) substrate
using the same deposition methodology.42 The interactions of the
copper substrate with the ceria nanoparticles are much stronger,
allowing the existence of small oxide nanoparticles on terraces of
the surface (CeOx-I) and forcing a unique morphology on the large
oxide particles (CeOx-II).

Figure 9.7 STM image taken after dosing Ce to Cu(111) at 650 K under an atmos-
phere of O2 (p ∼ 5 × 10−7 torr).16,17 Taken from Rodriguez et al.17. Copyright 2009 Wiley.

b1469_Ch-09.indd 477 4/8/2013 12:36:24 PM


b1469 Catalysis by Ceria and Related Materials

478 J. A. Rodriguez

CeO2 nanoparticles deposited on Rh(111),32 Pt(111),41 Cu(111),17


and Au(111)9 are much easier to reduce than extended surfaces of
bulk CeO2. Oxygen vacancies can be produced by annealing the
supported ceria nanoparticles in vacuum at elevated tempera-
tures32,41 or by direct reaction with CO or H2.9,17 The reducibility of
the ceria/metal systems can be an important property since bulk
ceria exhibits a low reactivity towards many molecules and is chemi-
cally active only after the generation of O vacancies.44 Thus, in the
ceria/metal systems one could couple the special reactivity of the
oxide nanoparticles to the reactivity of the metal to obtain highly
active catalysts.
Au(111) and Au(100) cannot dissociate water and, therefore,
are not able to catalyze the WGS reaction.9,12 Density functional cal-
culations have been used to compare the chemistry associated with
the WGS on Au(100) and Cu(100).12 On Cu(100), the first and the
most energy-consuming step, or rate-limiting step, is water dissocia-
tion with a ∆E of +9 kcal.mol−1 and a barrier of 26 kcal.mol−1. In con-
trast, the dissociation of adsorbed OH* and the subsequent
formation of CO2 and H2 are more facile.12 All the adsorbates bond
more weakly on Au(100) than on Cu(100).12 Consequently, the rate-
limiting dissociation of H2O on Au(100) is even more endothermic
(∆E = +17 kcal.mol−1) and the corresponding barrier is also higher
(35 kcal.mol−1). These theoretical results are in agreement with
experimental measurements, which show that Cu is a good WGS
catalyst while Au is an extremely poor one.9,28,45 The theoretical
results indicate that gold will be an excellent WGS catalyst if an
oxide helps with the dissociation of water.12 Once the OH groups are
formed, subsequent steps for the WGS process occur readily on the
gold substrate.9,12 Figure 9.8 displays XPS spectra obtained after
adsorbing water at 300 K on a ceria/gold system that contained fully
oxidized ceria nanoparticles (Ce4+) and on a system that contained
a substantial amount of Ce3+.46 The CeO2/Au(111) surface was not
able to dissociate water. In this respect, it behaved like a bulk
CeO2(111) surface.28 But the CeO2/Au(111) system can be easily
reduced with CO gas,9 and in the presence of Ce3+ cations there is a

b1469_Ch-09.indd 478 4/8/2013 12:36:24 PM


b1469 Catalysis by Ceria and Related Materials

Design and Modeling of Active Sites in Metal–Ceria Catalysts 479

Figure 9.8 (a) Typical Ce 3d core level spectra for Ce4+, from CeO2/Au(111), and
Ce3+, from Ce2O3/Au(111). (b) Two sets of O 1s core level spectra obtained from
CeO2/Au(111) and Ce2O3/Au(111) before and after dosing water at 300 K.46
Copyright 2007 Elsevier.

b1469_Ch-09.indd 479 4/8/2013 12:36:24 PM


b1469 Catalysis by Ceria and Related Materials

480 J. A. Rodriguez

Figure 9.9 Arrhenius plot for the WGS reaction rate on Cu(111),17 Cu(100),4 Cu/
ZnO(000ī),4 and on an Au(111) surface approximately 20% covered by ceria.15 The
data were acquired with a pressure of 20 torr of CO and 10 torr of H2O and tem-
peratures of 575, 600, 625, and 650 K.

fast dissociation of water.9,46 Eventually, the deposition of CeOx


nanoparticles on Au(111) produces excellent WGS catalysts9 as
shown in Fig. 9.9. This figure displays an Arrhenius plot for the WGS
activity of a CeOx/Au(111) surface in which 20% of the gold sub-
strate was covered by ceria.15 For comparison we also include results
obtained for the WGS on Cu(100),4 Cu(111),17 and Cu/ZnO(000ī)4
surfaces. The results in Fig. 9.9 indicate that the inverse CeOx/
Au(111) catalyst exhibits a larger WGS activity than those of copper
surfaces or even Cu nanoparticles dispersed on a ZnO(000ī) sub-
strate. On Cu(111) and Cu(100), the apparent activation energies
for the WGS are 18.1 and 15.2 kcal.mol−1, respectively.4,17 The appar-
ent activation energy decreases to 12.4 kcal.mol−1 on Cu/ZnO(000ī)4
and 10.3 kcal.mol−1 on CeOx/Au(111).15 In the inverse CeOx/
Au(111) catalyst, the reactants can interact with defect sites of
ceria nanoparticles, metal sites of the support, or the metal-oxide

b1469_Ch-09.indd 480 4/8/2013 12:36:25 PM


b1469 Catalysis by Ceria and Related Materials

Design and Modeling of Active Sites in Metal–Ceria Catalysts 481

interface.9,15 Thus, one can gain activity due to the active participa-
tion of the oxide in the catalytic reaction.9,15
Post-reaction characterization of the inverse CeOx/Au(111)
catalyst with XPS pointed to a Ce4+→Ce3+ transformation and identi-
fied a C 1s feature at 289–290 eV corresponding to either HCOO or
CO3 species on the CeOx/Au(111) surface.9 The mechanism for the
WGS in CeOx/Au(111) is assumed to be the following pathway:5,12

CO → CO (a)
H2O → H2O (a)
H2O (a) → OH (a) + H (a)
CO (a) +OH (a) → HCOx (a)
HCOx (a) → CO2 + H (a)
2H (a) → H2

A stable HCOx intermediate species must precede the formation


of H2 and CO2. It has been proposed that the key intermediate spe-
cies for the WGS reaction is either a formate (HCOO) or a carbonate
(CO3).1,5 Recent theoretical calculations also suggest the possibility
of a carboxylate (HOCO) intermediate.12,47 These species have differ-
ent coordination modes (see Fig. 9.10) and different lifetimes on
the surface of the catalyst.
Adsorption of HCOOH and CO2 has been used to create HCOO
and CO3 groups on CeOx/Au(111) surfaces.15 The orientation of
the formate on the ceria nanoparticles supported on Au(111) is
likely to be near normal to the surface as a bidentate species in a
chelating or bridge conformation. No preferred orientation was

H O H
O O
C C C
O O O O

Ce Ce Ce Ce Ce
Formate Carbonate Carboxylate

Figure 9.10 Different coordination modes for a formate (HCOO), carbonate


(CO3), and carboxylate (HOCO) on a nanoparticle or surface of ceria.

b1469_Ch-09.indd 481 4/8/2013 12:36:25 PM


b1469 Catalysis by Ceria and Related Materials

482 J. A. Rodriguez

observed for adsorbed carbonate.15 HCOOads appeared to have


greater stability than CO3,ads with desorption temperatures up to
600 K while CO3 only survived on the surface up to 300 K. Both spe-
cies could be valid intermediates for the WGS because the reaction
temperatures in Fig. 9.9 are elevated (575–650 K).15
For the CeOx/Au(111) catalysts, the presence of Ce+3 led to the
dissociation of H2O to give OH groups.15,47 Figure 9.11 displays O 1s
spectra collected after adsorbing H2O on a CeO1.75/Au(111) surface
at 100 K with subsequent annealing to the indicated temperatures.
These data indicate that OH species are stable on the surface up to
600 K and could interact with CO to yield weakly bound intermedi-
ates. When there is an abundance of Ce+4, the OH concentration is
diminished and the likely intermediates are carbonates.15 As the
surface defects are increased and the Ce+3/Ce+4 ratio grows, the OH
concentration also grows and both carbonate and formate species
are observed on the surface after dosing CO to H2O/CeOx/
Au(111).15 Data from X-ray absorption spectroscopy allows the

2.2
CeOx - Cu(111)
O1s
2.0
Intensity (arb. units)

1.8
OH
1.6

1.4
300 - 600 K

1.2

1.0
538 536 534 532 530 528
Binding energy (eV)

Figure 9.11 XPS spectra in the O 1s region for a dose of H2O to CeO1.75/Au(111)
followed by annealing steps as depicted from 150 to 600 K.15 Copyright 2010 Elsevier.

b1469_Ch-09.indd 482 4/8/2013 12:36:25 PM


b1469 Catalysis by Ceria and Related Materials

Design and Modeling of Active Sites in Metal–Ceria Catalysts 483

0.6

288 C K-edge XAS

0.5

o
65
301.5
Intensity (arbitrary units)

0.4 286.1
290.4

0.3

288.4
o
303.6 0
0.2 290.3

0.1

0.0
284 288 292 296 300 304 308 312
Photon energy (eV)

Figure 9.12 C K-edge from near-edge X-ray absorption fine structure (NEXAFS)
plots of CO (3 L) dosed to hydroxylated CeO1.75/Au(111) at 100 K. Both the normal
incidence (0°) angle and the grazing incidence (65°) spectra are shown in the
figure.15 Copyright 2010 Elsevier.

identification of the species produced by the reaction of surface OH


with CO.15 Figure 9.12 shows spectra taken after dosing 3 L of CO
onto a heavily hydroxylated CeOx/Au(111) surface that had a large
Ce+3 concentration. This set of C K-edge data is complex with many
contributions evident.15 The peaks that appear at 288 and 288.4 eV
share close resemblance to that of the π∗C=O resonance in grazing
and normal incidence for adsorbed formates. The 290.4 and 290.3
eV peaks are close to the π∗C–O resonance of carbonates in grazing
and normal incidences. The peak at 286.1 eV is due to CO adsorbed
on the surface. Weaker, broader contributions at 301.5 and 303.6 eV
are likely to be a mixture of σ features from both CO3,ads and formate
HCOOads species.17 The data, however, do not contain the features
expected for an adsorbed carboxylate (HOCO) intermediate.15

b1469_Ch-09.indd 483 4/8/2013 12:36:25 PM


b1469 Catalysis by Ceria and Related Materials

484 J. A. Rodriguez

A possible scenario for reaction of CO with OH in the presence


of O is:15

[Ce+4] 2CO + 0.5OHads + 2Oads → COads + CO3,ads + 0.5OHads


COads + CO3,ads + 0.5OHads → 2CO2 + 0.25H2 + 0.5Oads
[Ce+3/Ce+4] 2CO + OHads + Oads → 0.5CO3,ads + HCO2,ads + 0.5COads
0.5CO3,ads + HCO2,ads + 0.5COads → 2CO2 + 0.5H2

The ratio of OH:O(oxide) on the surface dictates the chemical


specificity of the intermediate that arises.15 The addition of ceria
nanoparticles to Au(111) is essential to generate an active WGS cata-
lyst and to increase the production and stability of key reaction
intermediates (OH, HCOO, and CO3). The Ce+3 is critical to OH
formation and a likely scenario may be that the OH could spill over
from the CeOx nanoparticles on to the Au(111) surface.9,15 The car-
boxylate (HOCO) intermediate seen in DF calculations12,47 is experi-
mentally elusive and difficult to detect.15 The low thermal stability of
the carboxylate makes it an ideal transient species for the WGS12,47
and it may only be observed under steady-state conditions.
Cu(111) has become a benchmark surface for studying the WGS
reaction.17,28,45,47,48 Specific rates, activation energies, and reaction
orders are consistent with data reported for ZnO-supported Cu cata-
lysts.28,45 As in the case of Au(111), the rate-limiting step for WGS on
Cu(111) is the dissociation of water.17,28,47,49 DF calculations indicate
that the reaction is endothermic (∆E= 5–14 kcal.mol−1) with a large
activation energy (Ea = 20–32 kcal.mol−1).47,49 The deposition of ceria
nanoparticles on Cu(111) transforms the dissociation of water into
an exothermic process (∆E= -8 kcal.mol−1) with an activation energy
smaller than 8 kcal.mol−1.17 The CeOx nanoparticles supported on
Cu(111) generated a catalyst which was 8 (at 650 K) to 23 times (at
575 K) more active than the clean Cu(111) substrate. The WGS on
Cu(111) has an apparent activation energy of 18 kcal.mol−1, which is
2–5 kcal.mol−1 larger than that found on Cu(100) and Cu(110).17,28,45
In contrast, the apparent activation energy for the WGS on CeOx/
Cu(111) is only 7 kcal.mol−1. This value is also smaller than the
corresponding value found for Cu nanoparticles supported on

b1469_Ch-09.indd 484 4/8/2013 12:36:25 PM


b1469 Catalysis by Ceria and Related Materials

Design and Modeling of Active Sites in Metal–Ceria Catalysts 485

CeO2(111), 9 kcal.mol−1.4 Thus, CeOx/Cu seems to be the best cata-


lyst in terms of activity or apparent activation energy. This is a con-
sequence of the special chemical properties of the ceria nanoparticles
and the configuration of the catalyst. Post-reaction surface analysis
indicated that the dominant oxidation state of cerium in the CeOx/
Cu(111) catalyst was “+3.” This was not the case for the Cu/
CeO2(111) catalyst,4 where there were only a few O vacancies in the
ceria support.

9.5 Water Gas Shift Reaction on M/CeOx/TiO2(110)


Surfaces (M=Cu, Au, or Pt)
The studies described above indicate that highly active WGS cata-
lysts are bifunctional with the metal and oxide catalyzing different
parts of the reaction. To optimize the performance of these systems
one must enhance the participation of the metal and oxide phases
in the catalytic process. Figure 9.3 shows three different configura-
tions in which a metal and an oxide can be combined in a catalyst.
In a conventional metal/oxide configuration, one enhances the
reactivity of the metal but this usually covers the defect sites of the
oxide (nucleation centers for the metal particles), which have chem-
ical activity. This is not the case in the inverse oxide/metal catalyst,
which enhances the participation of the oxide in the catalytic reac-
tion.16 In the inverse catalyst, the reactants can interact with defect
sites of oxide nanoparticles, metal sites of the support, or the metal/
oxide interface.16 In the quest to optimize the reactivity of the metal
and oxide phases, the most complex and promising configuration is
a mixed metal oxide array in which nanoparticles of a metal and an
oxide can interact with the reactants. Extremely active WGS catalysts
have been found after coadsorbing nanoparticles of ceria and
metals (Au, Cu, or Pt) on a TiO2(110) substrate.18,50
In recent years there has been strong interest in obtaining a
fundamental understanding of the behavior of cerium cations in
mixed metal oxide systems.43,51 In principle, the combination of
two metals in an oxide matrix can produce materials with novel
structural or electronic properties that can lead to superior catalytic

b1469_Ch-09.indd 485 4/8/2013 12:36:25 PM


b1469 Catalysis by Ceria and Related Materials

486 J. A. Rodriguez

Figure 9.13 (a) STM image of CeOx on a TiO2(110) surface after depositing Ce
atoms at 600 K in O2 (PO = 1 × 10−7 torr) and subsequent annealing at 900 K in O2
2
(PO2 = 1 × 10−4 torr, Vt = 1.2 V, It = 0.07 nA). (b) Bias-dependent STM images of a
diagonal array of CeOx nanoparticles taken at the imaging bias of 1.2 V, 0.06 nA
(top) and 0.4 V, 0.06 nA (bottom).18,50

activity or selectivity.1,51–54 The deposition of Ce atoms or CeOx clus-


ters on a TiO2(110) substrate produces structures that are very
different from those of bulk ceria or ceria nanoparticles supported
on metal surfaces or silica films.18,50,51 Figure 9.13(a) shows an STM
image obtained after depositing Ce on a TiO2(110) surface at 600 K
in O2 (∼ 1 × 10−7 torr) and subsequent annealing at 900 K still under
an O2 environment.18 The image was taken after cooling the sample
down to 300 K and removing the O2 from the background. Two dif-
ferent types of features are observed on the terraces: diagonal arrays
of small bright spots (see solid green arrow) and bigger ones (see
dashed green arrow). The bigger clusters in Fig. 9.13(a) can be cor-
related with TiOx islands (∼ 3 Å in height), which resulted from the
reaction of oxygen gas with interstitial Ti from the reduced bulk.18
The diagonal arrays (see solid arrow) have a distinctive height (∼ 1.4
Å) and correspond to CeOx nanoparticles.18 To better understand
the structures of the CeOx nanoparticles, a bias-dependent STM meas-
urement was performed as shown in Fig. 9.13(b). These two images
display the same diagonal arrays of CeOx nanoparticles, but obtained
at different imaging bias. The top STM image was taken with an
imaging bias of +1.2 V and the angle of the diagonal array is close to

b1469_Ch-09.indd 486 4/8/2013 12:36:25 PM


b1469 Catalysis by Ceria and Related Materials

Design and Modeling of Active Sites in Metal–Ceria Catalysts 487

42° with respect to the [1–1 0] direction. When this feature was
imaged at +0.4 V, the individual bright features appeared as dimers.18
Each ceria dimer is located in between two rows of oxygen species
protruding from the surface. Spectra from XPS and ultraviolet pho-
toelectron spectroscopy indicated that the oxidation state of the Ce
atoms in the dimers was essentially +3.18,50 The titanium cations were
mainly Ti4+ with a small amount (< 5 %) of Ti 3+ comparable to that
found on clean TiO2(110).18,50
DF calculations were used to study the bonding configuration of
the ceria dimers on TiO2(110), as shown in Fig. 9.14.50 A Ce atom
interacts with two bridging and one in-plane O atom of the titania
surface. Starting with this structure, the dissociation of an O2 mole-
cule near the adsorbed Ce is a highly exothermic process and an
energy of 6.88 eV is released.50 If a second Ce atom is deposited on
this surface, the CeO2 monomer would be an excellent binding site
to create a dimer. At the end, the ceria dimer adopts a configuration
comparable to STM images in Fig. 9.13(b) and has an oxidation
state of +3, without the generation of Ti3+ species, as seen in the XPS
measurements.50
The surface in Fig. 9.13(a) can be used as a model to study the
catalytic behavior of mixed metal oxide supports.10 CeOx nanoparti-
cles drastically affect the growth mode of Au, Cu, and Pt on
TiO2(110).10,40 On this surface, the admetals grow forming three-
dimensional particles. For example, gold exhibits very weak interac-
tions with the ideal terraces of TiO2(110) and mainly binds to
defects or step sites.24,55 Figure 9.15 shows STM images acquired
from the same surface area before (a) and after (b) depositing gold

Figure 9.14 Most stable geometries for the formation of a Ce2O3 dimer on
TiO2(110). DF calculations were used to study the interaction of Ce and O2 with the
titania substrate.10,50

b1469_Ch-09.indd 487 4/8/2013 12:36:25 PM


b1469 Catalysis by Ceria and Related Materials

488 J. A. Rodriguez

Figure 9.15 STM images for (a) CeOx/TiO2(110) and (b) Au/CeOx/TiO2(110).
The gold was deposited on the area shown in (a) at ∼ 300 K. ∼ 6% of the surface was
covered with Au. Taken in part from Park et al.18 Copyright 2009 National Academy
of Sciences USA.

on CeOx/TiO2(110) at ∼ 300 K.18 The CeOx/TiO2(110) was pre-


annealed under O2 (∼ 1 × 10−7 torr) at 900 K and had a morphology
similar to that seen in Fig. 9.13(a). The deposition of Au at room
temperature, ∼ 0.2 ML, produced three-dimensional metal particles
anchored to steps of the titania surface, “a” sites, to the (1 × 2) recon-
structions of TiO2(110), “b” sites, and to the CeOx dimers, “c” sites.18
On CeOx/TiO2(110), the dispersion of the Au was substantially
larger than seen on a pure TiO2(110) surface where Au mainly
binds to the steps.24,55
Neither CeOx/TiO2(110) nor TiO2(110) are able to catalyze
the WGS.18 However, Au/CeOx/TiO2(110) surfaces are outstand-
ing catalysts for the WGS as shown in Fig. 9.16. In test experiments,
Au/TiO2(110) surfaces were prepared following the same steps
used for the synthesis of Au/CeOx/TiO2(110) but without the
deposition of cerium.18 The Au/TiO2(110) systems were good WGS
catalysts, Fig. 9.16, but they did not come close to matching the
activity of Au/CeOx/TiO2(110).18 The same is valid when compar-
ing the WGS activities of Au/CeO2(111),4 CeOx/Au(111),9,15 Cu/
ZnO(000ī),4 and copper single crystals.4,21,37 The large dispersion
of gold on CeOx/TiO2(110), Fig. 9.15(b), should lead to a high
catalytic activity. The Ce3+ sites present in CeOx/TiO2(110) easily

b1469_Ch-09.indd 488 4/8/2013 12:36:26 PM


b1469 Catalysis by Ceria and Related Materials

Design and Modeling of Active Sites in Metal–Ceria Catalysts 489

-2
molec cm
20 torr CO, 10 torr H2O
20
625 K
17

15 on CeOx/TiO2(110)
Molecules produced / 10

10

on TiO2(110)
5

H2
CO2
0
0.0 0.4 0.8 1.2 1.6 2.0 2.4
Au coverage / ML

Figure 9.16 Water gas shift activity of Au/TiO2(110) and Au/CeOx/TiO2(110) as


a function of Au coverage.18 The area of TiO2(110) covered by CeOx was measured
with ion scattering spectroscopy (ISS), before depositing gold and found to be
∼ 12% of the clean substrate.18 The reported values for the production of H2 and
CO2 were obtained after exposing the catalysts to 20 torr of CO and 10 torr of H2O
at 625 K for 5 min.18 Copyright 2009 National Academy of Sciences USA.

dissociate water50 but, on exposure to CO, highly stable HCOx spe-


cies were formed on the oxide surface and there was no production
of H2 or CO2 gas.18,50 In Au/CeOx/TiO2(110), one has a bifunc-
tional catalyst: the adsorption and dissociation of water take place
on the oxide, CO adsorbs on the gold nanoparticles, and all subse-
quent reaction steps occur at oxide/metal interfaces. Au nanopar-
ticles do catalyze the reaction of OH with CO to yield a HOCO
intermediate and then H2 and CO2.56,57
The presence of ceria on TiO2(110) also improves the disper-
sion of Cu and Pt on the titania surface.50 Figure 9.17 shows the
amount of H2 produced by the WGS on catalysts generated by
depositing 0.15 ML of Cu, Pt, or Au on CeOx/TiO2(110) surfaces in
which 12–14% of the titania was pre-covered with CeOx nanoparti-
cles.50 For comparison, we also include the corresponding data for
the deposition of 0.15 ML of the admetals on plain TiO2(110). It is
clear that the nature of the oxide support strongly affects catalytic

b1469_Ch-09.indd 489 4/8/2013 12:36:26 PM


b1469 Catalysis by Ceria and Related Materials

490 J. A. Rodriguez

Pt nano

H2 produced / 1017 molecule cm-2


WGS
20
Cu nano
15

10
Au nano

0
TiO2 CeOx/TiO2 TiO2 CeOx/TiO2 TiO2 CeOx/TiO2

Figure 9.17 Amount of H2 produced by the WGS reaction on catalysts generated by


depositing 0.15 ML of Au, Cu, or Pt on CeOx/TiO2(110) surfaces in which 12–14%
of the titania was pre-covered with CeOx nanoparticles.50 For comparison, we also
include data for the deposition of the metals on TiO2(110). The reported values for
the production of H2 were obtained after exposing the catalysts to 20 torr of CO and
10 torr of H2O at 625 K for 5 min. The number of H2 molecules produced is normal-
ized by the sample surface area.50 Copyright 2010 the American Chemical Society.

activity. The catalysts containing CeOx/TiO2(110) display an activity


that is 2.5–3.8 times higher than those containing plain TiO2(110)
or CeO2(111).4 When compared to Cu(111), a typical benchmark
in studies of the WGS,17,28,45,47,48 or Cu/ZnO, a common WGS cata-
lyst for industrial applications,1 the M/CeOx/TiO2(110) (M = Cu,
Au, Pt) systems are 5–15 times more active. Catalytic activity
increases following the sequence: Au/CeOx/TiO2(110) < Cu/CeOx/
TiO2(110) < Pt/CeOx/ TiO2(110). Although Au is the admetal with
the lowest catalytic activity in Fig. 9.17, it is the one most affected by
coupling to CeOx/TiO2(110), because neither extended surfaces of
Au nor Au nanoparticles are able to catalyze the WGS.9 The behav-
ior of Cu/CeOx/TiO2(110) is remarkable. Supported copper
exhibits a WGS activity somewhat smaller than that of Pt, but it is
much less expensive.
In M/CeOx/TiO2(110) systems, there is a strong coupling of the
chemical properties of the admetal and the mixed metal oxide.18,50

b1469_Ch-09.indd 490 4/8/2013 12:36:26 PM


b1469 Catalysis by Ceria and Related Materials

Design and Modeling of Active Sites in Metal–Ceria Catalysts 491

The high catalytic activity of M/CeOx/TiO2(110) can be attributed


to the special chemical properties of the Ce2O3 dimers supported on
titania. Usually, the rate-determining step in the WGS reaction is the
dissociation of water.12,47 The Ce3+ sites present in CeOx/TiO2(110)
easily dissociate water.50 Thus, the M/CeOx/TiO2(110) systems are
bifunctional catalysts: adsorption and dissociation of water takes
place on the oxide, CO adsorbs on metal nanoparticles, and all sub-
sequent reaction steps occur at oxide/metal interfaces.50 The results
in Fig. 9.17 illustrate the tremendous impact that an optimization of
the chemical properties of nanoceria can have on the activity of a
WGS catalyst. This has been verified in experiments with a high-
surface-area powder Pt/CeOx/TiO2 catalyst, which in many aspects
shows behavior similar to that seen for Pt/CeOx/TiO2(110) sur-
faces.19 The approach should be valid in general for catalysts that
contain ceria as part of a mixed metal oxide, opening new directions
for tuning catalytic activity by coupling appropriate pairs of oxides.
The key issue is to take advantage of the complex interactions that
occur in a mixed metal oxide at a nanometer level.50,51

9.6 Conclusion
Experiments performed using model catalysts indicate that ceria
can play a key role in the performance of water gas catalysts.
Although zinc oxide is frequently used in industrial WGS catalysts,
the Cu/ZnO(000ī) and Au/ZnO(000ī) systems display low WGS
activity compared to CuCeO2(111) and Au/CeO2(111). A ceria sup-
port contains a substantial number of metal cations that are not fully
oxidized under WGS reaction conditions and may participate
directly in the dissociation of water and other important steps of the
catalytic process. Extremely active WGS catalysts have been found
after coadsorbing nanoparticles of ceria and noble metals on a
TiO2(110) substrate. The M/CeOx/TiO2(110) (M=Cu, Au, or Pt)
system takes advantage of the complex interactions that occur in a
mixed metal oxide at the nanometer level, opening new directions
for tuning catalytic activity by coupling appropriate pairs of oxides.

b1469_Ch-09.indd 491 4/8/2013 12:36:26 PM


b1469 Catalysis by Ceria and Related Materials

492 J. A. Rodriguez

Acknowledgements
Many of the experiments described in this chapter were done by the
Group of Catalysis and Surface Science at Brookhaven National
Laboratory (L. Barrio, M. Estrella, J.C. Hanson, J. Hrbek, P. Liu,
S. Ma, A. Nambu, J.-B. Park, J.A. Rodriguez, S.D. Senanayake,
D. Stacchiola, X. Wang, W. Wen) in collaboration with research
groups at the Universidad Central de Venezuela (J. Evans, M. Pèrez)
and the Universidad de Sevilla (J. Graciani, J.F. Sanz). The research
carried out at Brookhaven National Laboratory was supported by
the US Department of Energy (Chemical Sciences Division,
DE-AC02–98CH10886).

References
1.Thomas, J.M., Thomas, W.J., Principles and Practice of Heterogeneous
Catalysis, VCH, New York, (1997).
2. Si, R., Flytzani-Stephanopoulos, M., Angew. Chem. Int. Ed., 47 (2008)
2884–2889.
3. Fu, Q., Saltsburg, H., Flytzani-Stephanopoulos, M. Science, 301 (2003)
935–938.
4. Rodriguez, J.A., Liu, P., Hrbek, J., Evans, J., Pérez, M., Angew. Chem.
Int. Ed., 46 (2007) 1329–1332.
5. Burch, R., Phys. Chem. Chem. Phys., 8 (2006) 5483–5500.
6. Rodriguez, J.A., Pérez, M., Evans, J., Liu, G., Hrbek, J., J. Chem. Phys.,
122 (2005) 241101.
7. Rodriguez, J.A., Pérez, M., Jirsak, T., Evans, J., Hrbek, J., González, L.,
Chem. Phys. Lett., 378 (2003) 526–530.
8. Fu, Q., Deng, W., Saltsburg, M. Flytzani-Stephanopoulos, M., Appl.
Catal. B: Environ., 56 (2005) 57–68.
9. Rodriguez, J.A., Ma, S., Liu, P., Hrbek, J., Evans, J., Pérez, M., Science,
18 (2007) 1757–1760.
10. Ricote, S., Jacobs, G., Milling, M., Ji, Y.Y., Patterson, P.M., Davis, B.H.,
Appl. Catal. A: Gen., 303 (2006) 35–41.
11. Bunluesin, T., Gorte, R.J., Graham, G.W., Appl. Catal. B: Environ., 15
(1998) 107–112.

b1469_Ch-09.indd 492 4/8/2013 12:36:26 PM


b1469 Catalysis by Ceria and Related Materials

Design and Modeling of Active Sites in Metal–Ceria Catalysts 493

12. Liu, P., Rodriguez, J.A., J. Chem. Phys., 126 (2007) 164705.
13. Wang, X., Rodriguez, J.A., Hanson, J.C., Gamarra, D., Martinez-Arias, A.,
Fernandez-Garcia, M., Top. Catal., 49 (2008) 81–88.
14. Senanayake, S.D., Stacchiola, D., Liu, P., Mullins, C.B., Hrbek, J.,
Rodriguez, J.A., J. Phys. Chem. C, 113 (2009) 19536–19544.
15. Senanayake, S.D., Stacchiola, D., Evans, J., Estrella, M., Barrio-Pliego, L.,
Pérez, M., Hrbek, J., Rodríguez, J.A., J. Catal., 271 (2010) 392–401.
16. Rodriguez, J.A., Hrbek, J., Surf. Sci., 604 (2010) 241–245.
17. Rodriguez, J.A., Graciani, J., Evans, J., Park, J.B., Yang, F., Stacchiola, D.,
Senanayake, S.D., Ma, S., Perez, M., Liu, P.J., Sanz, J.F., Hrbek, J.J.,
Angew. Chem. Int. Ed., 48 (2009) 8047–8050.
18. Park, J.B., Graciani, J., Evans, J., Stacchiola, D., Ma, S., Liu, P., Nambu, A.,
Sanz, J.F., Hrbek, J., Rodriguez, J.A. Proc. Natl. Acad. Sci. (PNAS), 106
(2009) 4975–4980.
19. Gonzalez, I.D., Navarro, R.M., Wen, W., Marinkovic, N., Rodríguez, J.A.,
Rosa, F., Fierro, J.L.G., Catal. Today, 149 (2010) 372–380.
20. Wang, X., Rodriguez, J.A., Hanson, J.C., Gamarra, D., Martinez-Arias, A.,
Fernandez-Garcia, M., J. Phys. Chem. B, 110 (2006) 428–434.
21. Wang, X., Rodriguez, J.A., Hanson, J.C., Gamarra, D., Martinez-Arias, A.,
Fernandez-Garcia, M., J. Phys. Chem. B, 109 (2005) 19595–19603.
22. Wang, X., Rodriguez, J.A., Hanson, J.C., Perez, M., Evans, J., J. Chem.
Phys., 123 (2005) 221101.
23. Walden, M., Lai, X., Goodman, D.W., Science, 281 (1998) 1647–1651.
24. Park, J.B., Ratliff, J.S., Ma, S., Chen, D.A., J. Phys. Chem. C, 111 (2007)
2165–2171.
25. Baron, M., Bondarchuk, O., Stacchiola, D., Shaikhutdinov, S., Freund,
H.J., J. Phys. Chem. C, 113 (2009) 6042–6049.
26. Rodríguez, J.A., Evans, J., Graciani, J., Park, J.-B., Liu, P., Hrbek, J.,
Sanz, J.F., J. Phys. Chem. C, 113 (2009) 7364–7370.
27. Rodriguez, J.A., Liu, P., Hrbek, J., Pérez, M., Evans, J., J. Molecular
Catal. A: Chem., 281 (2008) 59–65.
28. Campbell, C.T., Koel, B., Daube, K.A., J. Vac. Sci. Technol. A, 5 (1987)
810–815.
29. Mullins, D.R., Radulovic, P.V., Overbury, S.H., Surf. Sci., 429 (1999)
186–197.

b1469_Ch-09.indd 493 4/8/2013 12:36:26 PM


b1469 Catalysis by Ceria and Related Materials

494 J. A. Rodriguez

30. Matolin, V., Sedlacek, L., Matolinova, I., Sutara, F., Skala, T., Smid, B.,
Libra, J., Nehasil, V., Prince, K.C., J. Phys. Chem. C, 112 (2008)
3751–3758.
31. Staudt, T., Lykhach, Y., Hammer, L., Schneider, M.A., Matolin, V.,
Libuda, J., Surf. Sci., 603 (2009) 3382–3388.
32. Castellarin-Cudia, C., Surnev, S., Schneider G., Podlucky, R., Ramsey,
M.G., Netzer, F.P., Surf. Sci., 554 (2004) L120–124.
33. Mullins, D.R., Robbins, M.D., Zhou, J., Surf. Sci., 600 (2006)
1547–1551.
34. Lu, J.-L., Gao, H.-J., Shaikhutdinov, S., Freund, H.-J., Surf. Sci., 600
(2006) 5004–5010.
35. Alexandrou, M., Nix, R.M., Surf. Sci., 321 (1994) 47–57.
36. Xiao, W., Guo, Q., Wang, E.G., Chem. Phys. Lett., 368 (2003) 527–531.
37. Schierbaum, K.-D., Surf. Sci., 399 (1998) 29–34.
38. Berner, U., Schierbaum, K., Phys. Rev. B, 65 (2002) 235404.
39. Berner, U., Schierbaum, K., Johnes, G., Wincott, P., Haq, S., Thornton,
G., Surf. Sci., 467 (2000) 201–208.
40. Grinter, D., Pang, C.L., Ithnin, R., Thornton, G., J. Phys. Chem. C, 114
(2010) 17036–17041.
41. Hardacre, C., Roe, G.M., Lambert, R.M., Surf. Sci., 326 (1995) 1–10.
42. Ma, S., Rodriguez, J.A., Hrbek, J., Surf. Sci., 602 (2008) 3272–3279.
43. Eck, S., Castellarin-Cudia, C., Surnev, S., Ramsey, M.G., Netzer, F.P.,
Surf. Sci., 520 (2002) 173–185.
44. Trovarelli, A. Catal. Reviews, 38 (1996) 439–520.
45. Nakamura, J., Campbell, J.M., Campbell, C.T., J. Chem. Soc. Faraday
Trans., 86 (1990) 2725–2730.
46. Zhao, X., Ma, S., Hrbek, J., Rodriguez, J.A., Surf. Sci., 601 (2007)
2445–2452.
47. Gokhale, A.A., Dumesic, J.A., Mavrikakis, M., J. Am. Chem. Soc., 130
(2008) 1402–1409.
48. Callaghan, C.A., Vilekar, S.A., Fishtik, I., Datta, R., Appl. Catal. A, 345
(2008) 213–232.
49. Fajín, J.L.C., Illas, F., Gomes, J.R.B., J. Chem. Phys., 130 (2009) 224702.
50. Park, J.B., Graciani, J., Evans, J., Stacchiola, D., Senanayake, S.D.,
Barrio, L., Liu, P., Sanz, J.F., Hrbek J., Rodriguez, J.A., J. Am. Chem.
Soc., 132 (2010) 356–363.

b1469_Ch-09.indd 494 4/8/2013 12:36:26 PM


b1469 Catalysis by Ceria and Related Materials

Design and Modeling of Active Sites in Metal–Ceria Catalysts 495

51. Rodriguez, J.A., Stacchiola, D., Phys. Chem. Chem. Phys., 12 (2010)
9557–9565.
52. Kung, H.H., Transition Metal Oxides: Surface Chemistry and Catalysis,
Elsevier, Amsterdam, (1989).
53. Muylaert, I., Van Der Voort, P., Phys. Chem. Chem. Phys., 11 (2009)
2826–2832.
54. Rodriguez, J.A., Catal. Today, 85 (2003) 177–192.
55. Valden, M., Lai, X., Goodman, D.W., Science, 281 (1998) 1647–1651.
56. Ojifinni, R.A., Froemming, N.S., Gong, J., Pan, M., Kim, T.S.,
White, J.M., Henkelman, G., Mullins, C.B., J. Am. Chem. Soc., 130
(2008) 6801–6812.
57. Kim, T.S., Gong, J., Ojifinni, R.A., White, J.M., Mullins, C.B., J. Am.
Chem. Soc., 128 (2006) 6282–6283.

b1469_Ch-09.indd 495 4/8/2013 12:36:26 PM


b1469 Catalysis by Ceria and Related Materials

This page intentionally left blank

b1469_Ch-09.indd 496 4/8/2013 12:36:26 PM


b1469 Catalysis by Ceria and Related Materials

CHAPTER 10

CERIA-BASED GOLD CATALYSTS:


SYNTHESIS, PROPERTIES,
AND CATALYTIC PERFORMANCE
FOR THE WGS AND PROX PROCESSES
Donka Andreeva, Tatyana Tabakova
and Lyuba Ilieva
Institute of Catalysis, Bulgarian Academy of Sciences,
Acad. G. Bonchev St., Bldg.11, 1113 Sofia, Bulgaria

10.1 Introduction
During the past two decades catalysis by gold has been one of the
fastest growing fields in catalytic research.1 The discovery by Haruta
and co-workers that highly dispersed gold nanoparticles supported
on reducible oxide supports exhibit an exceptional activity in CO
oxidation below ambient temperature inspired an increased level of
interest in the catalytic studies of gold.2 Currently, there is an appre-
ciable amount of research into the application of gold catalysts for
promoting many reactions of both environmental and industrial
significance. In particular, the water gas shift (WGS) and preferential
CO oxidation (PROX) reactions are important process steps in one
very attractive approach to CO-free hydrogen production for fuel-
cell applications. In the past decade, research has demonstrated the
very high potential of fuel cells to replace the internal-combustion
engine in vehicles and to provide power for stationary and portable

497

b1469_Ch-10.indd 497 4/8/2013 12:37:28 PM


b1469 Catalysis by Ceria and Related Materials

498 D. Andreeva, T. Tabakova and L. Ilieva

power applications because they are energy efficient, clean, and fuel
flexible.3 Hydrogen has a great potential as an environmentally
clean, renewable, and highly efficient energy carrier. Nowadays,
hydrogen is mostly produced from fossil fuels by a multistep process
that includes catalytic reforming or autothermal reforming of hydro-
carbons, followed by the WGS reaction (CO + H2O ↔ CO2 + H2).
This reaction has a long historical application as an industrially
important process for hydrogen production. Recently it has attracted
renewed interest due to the increasing demands for high-purity
hydrogen. The hydrogen-rich gas stream after the WGSR contains
typically 0.5–1 vol% CO. Due to the high sensitivity of the platinum
anode electrode in polymer electrolyte membrane fuel cells
(PEMFCs) towards even low levels of CO, the concentration of CO
should be reduced to below 10 ppm (below 100 ppm for CO-tolerant
alloy anodes). The selective catalytic oxidation of CO in a hydrogen-
rich gas stream, known as the PROX reaction, appears to be the
simplest and most cost-effective method for this purpose.
The development of fuel-cell technology has focused consider-
able research interest on the design of active, stable, and poison-
resistant catalysts for a low-temperature WGSR as well as active and
highly selective catalysts for PROX. Supported gold catalysts have
been extensively studied during the last two decades in view of their
high WGS activity at low temperatures.4–10 The nature of the sup-
port on which nanosized gold particles are dispersed plays a crucial
role in determining catalytic activity. Recent investigations have
revealed the beneficial application of ceria as a support for active
catalysts for WGS and PROX reactions. The distinctive defect chem-
istry of ceria and its ability to exchange lattice oxygen with the gas
phase results in an oxide with unique catalytic properties, including
the promotion of precious-metal dispersion and the enhancement
of catalytic activity at interfacial metal-support sites.11 Numerous
studies of Au/CeO2 catalysts have addressed how various factors,
including the method of preparation, gold loading and support
surface area, gold particle size, structure and oxidation state, nature
of the active gold species, and reactivity of surface species, correlate
with catalytic performance in the WGSR.12–33

b1469_Ch-10.indd 498 4/8/2013 12:37:29 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Gold Catalysts 499

Gold–ceria catalysts have recently been proposed as promising


for the PROX reaction, too. Significant efforts have focused on dif-
ferent preparation methods and on the characterization and testing
of Au/CeO2 catalysts for the PROX reaction.34–44
Some researchers have reported the use of doped ceria or ceria
mixed oxides in order to improve the activity, selectivity, and stability
of gold catalysts for the WGS and PROX reactions.45–59 The modifica-
tion of ceria by divalent or trivalent ions affects its redox properties
and increases the oxygen vacancies in the ceria structure, and hence
increases the oxygen mobility and capacity of ceria-based catalysts.
The use of different dopants and different preparation methods in
the synthesis of ceria-based materials as supports for gold catalysts
can strongly affect catalytic performance.
Recent advances in the preparation of ceria-based gold catalysts
for hydrogen production by the WGS and PROX reactions are
reviewed in this chapter. Considerable emphasis is placed on the cata-
lyst characterization by a number of physicochemical methods: X-ray
diffraction (XRD), high-resolution transmission electron microscopy
(HRTEM), temperature programmed reduction (TPR), Raman spec-
troscopy, X-ray photoelectron spectroscopy (XPS), and Fourier trans-
form infrared (FTIR) spectroscopy. The relation between the structure,
properties, and catalytic activity, as well as the nature of the active sites
is also discussed.

10.2 Preparation of Gold–Ceria Catalysts


10.2.1 Preparation of ceria and doped ceria as supports
for gold catalysts
A comprehensive review of conventional processes for the synthesis
and modification of ceria-based materials was produced by Adachi
and Masui.60 Here we will describe some of the approaches that have
been recently employed to prepare ceria and ceria-based solids suited
for the synthesis of gold catalysts for the WGS and PROX reactions.
The application of ceria to support a gold catalyst requires
a number of specific features such as nano-scaled particle size,
crystalline structure, particular morphology, high surface area,

b1469_Ch-10.indd 499 4/8/2013 12:37:29 PM


b1469 Catalysis by Ceria and Related Materials

500 D. Andreeva, T. Tabakova and L. Ilieva

and thermal stability. One of the most often reported methods


for the preparation of powdered cerium oxide is homogeneous pre-
cipitation. Usually, the cerium salt Ce(NO3)3 6H2O is used. Different
precipitants are added: K2CO3,12,13 (NH4)2CO3,14 Na2CO3,16 or KOH.61
The reaction of the cerium salt under constant pH (usually 8–9)
with precipitants results in the precipitation of cerium hydroxide.
Thermal treatment above 300 °C leads to the formation of a nano-
sized well-crystallized material with a cubic fluorite structure and
average particle size 6–10 nm. The BET surface area is in the range
70–120 m2 g−1. The influence of different precipitants on the prop-
erties of ceria has not been studied systematically.
Flytzani-Stephanopoulos and co-workers prepared doped and
undoped ceria by the urea co-precipitation–gelation method and
found that their ceria-based samples had better homogeneity
and higher surface area than those prepared by conventional co-
precipitation with ammonium carbonate.14 In this case, the prepara-
tion procedure consists of mixing aqueous metal nitrate solutions
with urea (H2N–CO–NH2), heating the solution to 100 °C under
vigorous stirring and the addition of de-ionized water, boiling the
resulting gel for 8 h at 100 °C; the subsequent steps are filtering,
drying, and calcination.
Recent studies of ceria systems have focused on the develop-
ment of synthetic approaches to make size-controlled and shape-
controlled nanostructures. The thermolysis of an acidified Ce(NO3)4
solution was reported as a method for the preparation of nanocrys-
talline CeO2 (~ 4 nm, surface area 180 m2 g−1).39 The deposition of
gold on these nanocrystalline particles provides experimental evi-
dence for the size effect of supports. It was shown that the activity
of the resulting gold catalyst in CO oxidation was two orders of
magnitude higher than that of catalysts prepared by co-precipitation
or by Au deposition on a regular cerium oxide support (surface
area 70 m2 g−1).
Theoretical calculations of ceria surface structure indicate that
the (110) and (100) planes are more active than the (111) face.
Numerous studies have shown that ceria nanoparticles predomi-
nantly expose their most stable and therefore less reactive (111)

b1469_Ch-10.indd 500 4/8/2013 12:37:29 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Gold Catalysts 501

planes, and the generation of more reactive planes with defined


structures is difficult by conventional methods. Ceria nanocrystals in
the shape of rods, cubes, and polyhedra have been prepared by the
hydrothermal method.43 The mixing of aqueous solutions of
Ce(NO3)3 6H2O and different amounts of NaOH in a Teflon bottle
and stirring formed a milky slurry. Subsequently, the Teflon bottle
with this mixture was placed into a stainless steel autoclave inside a
temperature-controlled electric oven, and was subjected to hydro-
thermal treatment at temperatures in the range 100–180 °C for 24 h.
The CeO2 nanopolyhedra, nanorods, and nanocubes were of the
pure cubic phase with lattice constants of 5.414(3), 5.436(3), and
5.405(3) Å, respectively. According to transmission electron micros-
copy (TEM) images, the nanorods had a uniform width of approxi-
mately 10 nm with length 50–200 nm, the nanocubes showed an
average size of about 26 nm, and the nanopolyhedra displayed a
diameter around 10 nm.
Modification of the hydrothermal method of surfactant-assisted
synthesis allows control of the texture, structure, and morphology of
cerium oxide nanomaterials. Mesoporous structures and 1D
nanorods of CeO2 have been produced by Yuan et al.62 Synthesis was
performed with CeCl3 7H2O as the cerium oxide precursor in the
presence of the surfactant cetyltrimethylammonium bromide
(CTMABr). It was found that the surfactant concentration and stir-
ring during the hydrothermal treatment are key factors in influenc-
ing the final structure of the cerium oxide. The crystallite sizes of
CeO2 synthesized using a Ce/surfactant ratio of 1.2 and 0.33 by
static autoclaving, and synthesized using a Ce/surfactant ratio of 1.2
while stirring, were about 8.9, 11.4, and 6.4 nm, respectively. After
calcination in air for 6 h at 550 °C the nanorods had a fluorite single-
crystalline structure with lengths in the range 150–300 nm and diam-
eters ranging from 10 to 25 nm.
Ceria with a surface area of 203 m2 g−1 was synthesized via
decomposition of cerium carbonate Ce2(CO3)3·xH2O by Kim and
Thompson.18 The cerium carbonate was hydrolyzed in a mixture of
75% steam in N2 at 150 °C and then calcined in dry air at 400 °C
for 4 h.

b1469_Ch-10.indd 501 4/8/2013 12:37:29 PM


b1469 Catalysis by Ceria and Related Materials

502 D. Andreeva, T. Tabakova and L. Ilieva

Solution combustion synthesis (SCS) is becoming an attractive


technique for the synthesis of different advanced porous ceramic or
metallic materials. Combustion synthesis has several advantages over
conventional preparation methods including low processing costs
and a simple solvent-free and quick preparation route, leading to
homogeneous, high surface area materials. Lenite et al. prepared
ceria using SCS followed by calcination in calm air at 650 °C for 3 h
to obtain a crystalline structure.63
Considerable efforts have been spent on the preparation of
ceria-based mixed oxides or the modification of ceria by doping
with suitable additives. Supports of ceria–alumina (10 and 20 wt%
Al2O3) have been prepared by co-precipitation of Ce(NO3)3 6H2O
and Al(NO3)3 9H2O with a solution of K2CO3.20 The addition of
alumina has two purposes: to assist in the formation of oxygen
vacancies and to increase the thermal stability of the catalysts.
Results reveal that modification with alumina influences the average
size of ceria particles and leads to higher dispersion compared to
pure ceria. Catalyst stability was improved because alumina prevents
agglomeration of gold and ceria. Moreover, the addition of alumina
by co-precipitation increases the number of oxygen vacancies in the
ceria. However, the vacancies are located predominantly in the bulk
of ceria, thus being inaccessible for re-oxidation and resulting in
decreased WGS activity of the Au/CeO2–Al2O3. Mixed ceria–alumina
oxides have also been prepared by washcoating and by mechano-
chemical activation (MA).21 A typical washcoating process involves
the precipitation of cerium hydroxide using commercial alumina
with a high surface area, which is suspended in water using ultra-
sound. An aqueous solution of Ce(NO3)3 6H2O and K2CO3 is used
under vigorous stirring at constant pH 9.0 and temperature 60 °C.
After aging, washing, and drying, the solids are calcined in air at
400 °C for 2 h. MA has been carried out by mechanically mixing alu-
mina and vacuum-dried cerium hydroxide. Cerium hydroxide has
been prepared by precipitation as described above. The mixture was
subjected to MA by milling and further calcination. Comparing the
effect of the preparation method on the WGS activity of gold cata-
lysts supported on mixed ceria–alumina oxides, it can be concluded

b1469_Ch-10.indd 502 4/8/2013 12:37:29 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Gold Catalysts 503

that mechanical milling is a very suitable method for the prepara-


tion of mixed ceria-based oxides.64 This procedure creates powders
with a small particle size of a few nanometers with a high concentra-
tion of lattice defects. The application of MA in the synthesis of
ceria–alumina mixed oxides has increased the number of oxygen
vacancies on the surface of gold catalysts and the re-oxidation tem-
peratures are even lower than those of gold–ceria.64
Zhang et al. prepared ceria doped with different amounts of La
(1.5–12 at%) by co-precipitation of an aqueous solution of cerium
and lanthanum nitrates with K2CO3.65 La doping can improve the
activity of Au–ceria catalysts by stabilizing the ceria and modifying its
morphology. The surface area increased gradually, whilst there was a
trend for a slight decrease in ceria particle size and cumulative pore
volume with an increase in La loading. The highest WGS activity was
attained over a gold catalyst supported on ceria doped with 5 at% La.
Hydrolysis of cerium and lanthanum isopropoxide in a reverse
(water-in-oil) microemulsion has been used to synthesize highly
homogeneous mixed La/Ce-oxides (LaxCe1−xO2−0.5x) with various La
contents.55 The microemulsion was composed of water, isooctane,
poly(ethylene oxide)-block-poly(propylene oxide)-block-poly(ethylene
oxide) with 10 wt% ethylene oxide and 1-pentanol as a co-surfactant.
The metal isopropoxide precursor solution (Ce- and La-isopropoxide
in isopropanol) was added to the microemulsion at the desired
La:Ce ratio. After aging while continually stirring at room tempera-
ture (RT) and phase separation, the gel was washed, freeze-dried,
and finally calcined at 450 °C. A slight decrease in particle size was
found after La doping, from mean particle diameters of 7.6 nm for
CeO2 to 5.5 nm for La0.5Ce0.5O1.75. The reducibility of the mixed
oxides showed an optimum at about 25% La content, which could be
related to improved oxygen mobility due to the formation of oxygen
vacancies. The results evidenced that carefully controlled synthesis of
nanostructured catalysts with a uniform, tailored composition can be
used to control the reactive properties of these materials.
Washcoating was used in the preparation of CeO2–TiO2 mixed
oxides with two different ceria loadings (20 and 50 wt%).66 HRTEM
measurements showed that the deposition of gold caused atomically

b1469_Ch-10.indd 503 4/8/2013 12:37:29 PM


b1469 Catalysis by Ceria and Related Materials

504 D. Andreeva, T. Tabakova and L. Ilieva

dispersed gold in the regions of the supports, which is rich in defective


ceria, and the presence of small gold particles in the titania-rich
regions. FTIR spectroscopy measurements evidenced two different
kinds of gold sites: small clusters and ultra-small gold particles. Special
sites where CO is irreversibly bonded to the outgassing at RT, ascribed
to the small cationic clusters, were observed on both samples of the
oxidized Au/CeO2–TiO2 catalyst, characterized by a band at 2166 cm−1.
After reductive pretreatment, negatively charged clusters or flattened,
thin gold particles, characterized by a band at 2084 cm−1, were found.
The relative amount and the nature of the exposed sites on the two
samples were strongly dependent on the support composition, which
resulted in a different performance for WGS and PROX reactions.67
Promising supports for the gold catalysts were prepared by using ceria
(20 wt%) as an additive to mesoporous titania68 or meso-macroporous
binary metal oxides (TiO2–ZrO2).69 A high degree of synergistic inter-
action between the ceria and the support and a positive effect on the
structural and catalytic properties of the WGSR were reported. The
WGS activity tests demonstrated better performance for the gold–
ceria-modified mesoporous titania catalysts compared to gold catalysts
supported on CeO2 and mesoporous TiO2, as well as when compared
to a reference catalyst Au/TiO2 type A (World Gold Council).68
Structural characterization revealed the beneficial role of modifica-
tion with ceria in decreasing the degree of crystallinity of meso-
macroporous binary metal oxides and creating smaller particles.69 The
addition of ceria induced a significant improvement in the reducibility
of the meso-macroporous support, which enhanced WGS activity.
A modification of the pseudo-sol–gel method based on the ther-
mal decomposition of metallic propionates, was used to synthesize
ceria oxides doped with 10 mol% of Zr, Zn, and Fe.56 Briefly, calcu-
lated amounts of Ce(III) acetate, Zr(IV), Zn(II), and Fe(III) acetyl
acetonates were dissolved in propionic acid. After mixing by stir-
ring, the propionic acid excess was evaporated until a kind of resin
was obtained that was calcined at 500 °C for 2 h. The formation of a
solid solution was observed after doping with Zr and Fe, and there
was ZnO surface segregation with Zn. It is interesting to consider
the differing effects of the dopants. Doping with Zr enhances the

b1469_Ch-10.indd 504 4/8/2013 12:37:29 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Gold Catalysts 505

formation of surface oxygen vacancies, after addition of Zn there is


the same number of vacancies as pure CeO2, and Fe reduces the
number of oxygen vacancies. Nevertheless, an enhancement of
reducibility was registered for all of the doped ceria materials and a
high dispersion of the gold nanoparticles was achieved in all cases.
The results revealed a strong relation between the enhancement of
the reducibility and the increase in the CO conversion during
PROX reactions. Iron-modified ceria supports with different con-
tents of Fe (10, 25, and 50 mol%) were synthesized by the same
method.57 The formation of a solid solution was reported after the
introduction of 10 mol% Fe into the ceria network, while iron oxide
segregation was detected above 25%. The highest CO conversion
per atom of Fe incorporated was attained over the solid with 10%
Fe content, where the synergy between Ce and Fe was a maximum.
A gold catalyst prepared with this material showed high activity and
selectivity with excellent stability during PROX reactions, and had a
higher CO conversion than the Au/CeO2 catalyst at temperatures
below 150 °C.
A synergistic catalytic effect for conversion and selectivity during
PROX reactions was reported for gold catalysts supported on vari-
ous ceria-containing mixed oxides. Ce–Fe mixed oxides were pre-
pared by urea gelation co-precipitation and used to support gold
catalysts.58 A support with composition 50 wt% CeO2 –50 wt% Fe2O3
appeared beneficial not only for nucleation and the growth of
highly dispersed gold particles (1–1.8 nm), but also for activation of
oxygen and its mobility. Moreover, the presence of Fe2O3 in the sup-
port composition improved resistance to deactivation by CO2. The
advantage of an MnO2–CeO2 composite oxide with various Mn/Ce
atomic ratios in comparison to single oxides when supporting gold
catalysts for PROX, was demonstrated by Chang et al.46 Mixed Mn–Ce
oxides were prepared by incipient-wetness impregnation using
Mn(NO3)3 as the Mn precursor and commercial CeO2 (BET surface
area 90 m2 g−1). The synthesis of mixed CeO2–Co3O4 catalysts with
a Ce/Co atomic ratio from 0.1 to 0.6 was carried out by sol–gel pre-
cipitation.48 Hydrolysis of Co(NO3)2 6H2O or Ce(NO3)36H2O in
ethanol under stirring produced alcosol. The precipitation was

b1469_Ch-10.indd 505 4/8/2013 12:37:29 PM


b1469 Catalysis by Ceria and Related Materials

506 D. Andreeva, T. Tabakova and L. Ilieva

achieved with a sodium hydroxide solution at ambient temperature.


Au/CeO2–Co3O4 exhibited much higher catalytic activity in PROX
reactions than Au/Co3O4 and Au/CeO2.

10.2.2 Deposition of gold


The synthesis of nanosized gold catalysts is highly sensitive to the
preparation technique, with a large variation in the size of the gold
particles and in their interaction with the support. Methods of pre-
paring gold catalysts can be classified into two groups. The first group
produces a well-mixed material through the simultaneous formation
of the support and the gold precursor. The synthesis procedures
most often reported for making gold catalysts on ceria and ceria-
based supports for PROX and WGS reactions are co-precipitation
(CP) and combustion. Co-precipitation occurs after mixing aqueous
solutions of HAuCl4 and a cerium salt with (NH4)2CO314 or Na2CO37,35,46
at 60–70 °C or RT, respectively, at a constant pH value of 8 (9–10,
respectively).35 After aging, the precipitate is filtered, washed, and
dried at 100–120 °C, and calcined at 400 °C. The combustion method
to synthesize a Au/CeO2 catalyst is carried out by mixing aqueous
solutions of (NH4)2Ce(NO3)6, HAuCl4, and urea (fuel) in a borosili-
cate dish, followed by autoignition with a cold flame in a muffle fur-
nace preheated at 400 °C.41
The second group is based on the deposition or adsorption of
Au compounds on a preliminary prepared support. Deposition-
precipitation (DP) is one of the most applicable methods for the
synthesis of active gold-based catalysts. In particular, this method
involves the deposition of hydroxide or hydrated oxide onto the
surface of the support because of the gradually raising pH value of
the solution in which the support is suspended.1 In the literature
there is evidence that not all the gold in the solution is precipitated
when some particular metal oxides are used to support the gold cata-
lysts. However, DP is effective for depositing gold onto a ceria sur-
face. The complete deposition onto ceria of the desired gold loading
using an exact amount of HAuCl4 solution has been confirmed by
atomic adsorption analysis (AAA), as reported in many papers.

b1469_Ch-10.indd 506 4/8/2013 12:37:29 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Gold Catalysts 507

Some modifications of this method have been reported. Typically,


the gold phase was deposited onto ceria, which is thoroughly sus-
pended in water, via chemical interaction between HAuCl4·3H2O
and a precipitant (Na2CO3, K2CO3, (NH4)2CO3, or NaOH) under
vigorous stirring, while keeping the pH constant (6.5–8 is reported
in different papers).10,12,14,18,20–27,32,34,36 The resulting precipitate mate-
rial was aged and then carefully washed with de-ionized water until
no Cl− ions could be detected. The washed material was then dried
under a vacuum and calcined in air at 300–400 °C. Some researchers
prepared aqueous solutions of the Au precursor by adjusting the pH
and increasing the temperature to 70–80 °C.56,57,61,63 Ceria powder
was then added and continually stirred. The next steps are the usual:
filtering, washing, drying, and calcination.
The influence of pH and the molarity of the gold precursor
solution during DP on WGS activity has been studied in detail.63
It was found that the most useful preparation conditions for Au/
CeO2 were pH = 8.5 and M = 0.2 × 10−3. HRTEM images of a cata-
lyst prepared under these conditions show small gold particles
(1.5–2 nm) and a homogeneous distribution on the carrier, features
that are fundamental for good catalytic activity. The increase of M
to 1 × 10−3 during catalyst preparation caused the formation of
slightly larger particles (average size 3 nm), which are mainly dis-
tributed on grain borders. Both lower pH (7) and higher pH (10)
values significantly influenced gold particle size and led to the
appearance of large particles (15–20 nm) due to the agglomeration
of smaller ones.
Strong shape and crystal-plane effects of nanoscale ceria on the
PROX43 and WGS70 activity of Au–CeO2 catalysts has been observed.
The deposition-precipitation of gold on different facets of ceria
nanorods, nanocubes, and nanopolyhedra led to the production of
gold catalysts with different catalytic behavior. According to theoreti-
cal studies, the formation energy of anion vacancies for different
ceria surfaces follows the order (110) < (100) < (111).71 These results
imply that the formation of oxygen vacancies is easier on CeO2
(110) planes. Ceria nanorods with (100)/(110)-dominant surfaces
should be better sites for anchoring and dispersing very fine gold

b1469_Ch-10.indd 507 4/8/2013 12:37:29 PM


b1469 Catalysis by Ceria and Related Materials

508 D. Andreeva, T. Tabakova and L. Ilieva

clusters. XPS and H2-TPR measurements evidenced that a much


higher fraction of strongly bound gold species is present on the sur-
face of ceria nanorods and polyhedra than on nanocubes.70
Additionally, the analysis of the microstrain in a ceria lattice using
XRD data also demonstrated that a nanorod sample was the most
defective, the polyhedron sample the next, while a nanocube sam-
ple was the least strained. All these observations are consistent with
the “light-off” WGS reaction profiles, which rank activity in the fol-
lowing way: Au/ceria rods > Au/ceria polyhedra >> Au/ceria cubes.70
The same activity order was found in CO-PROX reactions — the
best performance was shown by a gold catalyst on CeO2 nanorods in
comparison with Au/polyhedra and Au/cubes.43
In order to examine the efficiency of Au as a promoter for ceria-
based WGS catalysts, Jacobs et al. used vapor phase grafting to pre-
pare nanocrystalline Au clusters.19 A physical mixture of cerium
oxide (BET surface area 120 m2 g−1), preheated under a high
vacuum (1 × 10−7 torr) at 350 °C to drive water out of the pores, and
dimethyl(acetylacetonate gold III) as gold precursor were loaded
into a sample tube and attached to a high-vacuum line. The mixture
was heated in stages from 80 °C to 130 °C, held for 1 h at each tem-
perature, and finally cooled to room temperature. In order to attain
full decomposition of the Au acetylacetonate, a calcination tempera-
ture of 250 °C in O2 was selected, while a temperature of 175 °C in
H2 was chosen as the initial temperature for catalyst activation.
HRTEM images and energy-dispersive X-ray spectroscopy (EDX)
analysis showed the dependence of gold particle size on gold load-
ing: very small, well-dispersed Au crystallites were observed, mainly
in the 1–2 nm diameter range and 1–3 nm diameter range for 0.25%
Au and 0.5% Au samples, respectively. For a catalyst with 1% Au,
larger particles in the 5–8 nm range were also found (in addition to
the smaller crystallites), while for 2.5% Au/ceria and 5.0% Au/ceria
catalysts, a heterogeneous distribution was obtained. In addition
to 1–5 nm Au crystallites, there were also much larger particles
(10–20 nm for 2.5% Au and up to 30 nm at 5.0% Au loading).
Comparative studies for the effect of the preparation method
on the WGS or PROX activity of gold–ceria catalysts clearly

b1469_Ch-10.indd 508 4/8/2013 12:37:29 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Gold Catalysts 509

demonstrated the advantage of deposition-precipitation over other


synthesis procedures. The main reason is the creation of gold nano-
particles with a narrower size distribution, which are mainly local-
ized on the surface of the supports without embedding of the active
species in the bulk of the support. Tabakova et al. modified this
procedure. They deposited gold hydroxide on freshly prepared
Ce(OH)3, aged only in mother liquor at 60 °C for 1 h, instead of on
well crystallized ceria.72 Cerium (3+) ions act as a reducing agent
during aging, converting Au3+ to Au0 during preparation. The color
of the catalyst prepared by this modified deposition-precipitation
(MDP) became dark-gray immediately after precipitation of the
HAuCl4, which indicates the presence of metallic gold. HRTEM
measurements showed high crystallinity of the ceria in samples pre-
pared by both DP and MDP after calcination at 400 °C. However,
pronounced differences in the gold particle size of the samples were
observed. The presence of very highly dispersed gold clusters (d
about 1 nm) on the surface of the DP-prepared catalyst was evi-
denced by energy dispersive X-ray spectroscopy (EDS) analysis. In
contrast, only large gold particles (average d ~ 15 nm) with a lower
dispersion were observed in the sample prepared by MDP, because
the agglomeration of small metallic gold particles precipitated on
freshly prepared Ce(OH)3 takes place during the calcination of the
precursor. Additionally, XPS and FTIR measurements indicated
large differences in the availability of active gold sites at the surface,
which strongly influence both PROX42 and WGS activity.69
The effect of the preparation method (co-precipitation or dep-
osition-precipitation) on the performance of PROX reactions was
rationalized on the basis of sequential precipitation, which occurred
due to the different solubilities of the hydroxide species formed dur-
ing the CP stage.61 With Au/ceria, the gold hydroxide, with a much
lower solubility product (Ksp = 5 × 10−46), should be formed before
Ce(OH)3 with Ksp = 1.6 × 10−20, thus favoring the aggregation of gold
hydroxide in larger particles partially covered by the cerium hydrox-
ide, which forms later. In contrast, in the presence of preformed
ceria (the DP method) the extremely low solubility of gold hydrox-
ide causes the formation of a high number of nucleation centers,

b1469_Ch-10.indd 509 4/8/2013 12:37:29 PM


b1469 Catalysis by Ceria and Related Materials

510 D. Andreeva, T. Tabakova and L. Ilieva

which interact with the support, resulting in small gold particles well
dispersed on the surface.
An analysis of scanning transmission electron microscope
(STEM) micrographs and elemental maps of Au/ceria catalysts pre-
pared by co-precipitation and deposition-precipitation show a large
difference in the structure of the ceria.14 Although XRD analysis
evidenced the fluorite oxide-type diffraction pattern of CeO2 in
both cases, the sample prepared by CP had a needlelike and layered
bulk structure while the DP sample had a uniform spherical struc-
ture. Differences were also registered for gold particle size: they had
an average size of 8 nm in the CP sample while in the DP sample, the
gold particles were < 5 nm.
By comparing the structure and redox features of Au/CeO2 pre-
pared by different methods, Arena et al. found that dispersion is not
the only parameter determining the reactivity of the studied cata-
lysts. The effect of preparation method (DP, CP, incipient-wetness
(IW), combustion (CB)) was studied with a special emphasis on the
amount of residual chlorine.41 It was concluded that different syn-
thesis routes produce substantial variations in the amount of resid-
ual chlorine, which controls the strength of the Au3+–CeO2
interaction resulting in changes in the reducibility of the active
phase. The following activity scale for CO oxidation was observed:
DP > CP >> CB > IW. The removal of residual chlorine by washing in
an alkali solution enhances the reducibility of the active phase, lev-
eling off the functionality of the Au/CeO2 system in both total and
preferential CO oxidation reactions.
One of the main drawbacks of ceria-based catalysts is deactiva-
tion with time-on-stream. In order to overcome this Cargnello
et al. applied a new method to create very small Au nanoparticles
(average diameter > 2 nm) embedded inside the CeO2 support.
There was good thermal stability after aging treatments at
relatively high temperatures (up to 250 °C).44 The aim of this
methodology was to stabilize the Au phase by encapsulation of
preformed Au nanoparticles inside a porous ceria layer. A mixture
of thiolates was used to protect the gold particles and to direct
the self-assembly assisted precipitation of Ce(OH)x species around

b1469_Ch-10.indd 510 4/8/2013 12:37:29 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Gold Catalysts 511

the preformed particles. The removal of the organic layer was car-
ried out by calcination. Although very good catalytic activity was
obtained with low Au loading (1 wt%) under real PROX condi-
tions, CO chemisorption experiments demonstrated lower acces-
sibility of the gold phase over embedded catalysts in comparison
with reference Au/ceria samples prepared by DP. TPR experi-
ments on the embedded samples also suggested that a fraction of
the Au nanoparticles was completely buried inside the ceria layer
and further improvements could be achieved by better tuning the
porosity of the support.

10.3 Characterization of Gold–Ceria Catalysts


10.3.1 Dispersion and oxidation state of supported gold
The characterization of gold focuses on gold dispersion and the
oxidation state. It is considered that not only the dispersion but also
the morphology of the gold particles is important for rationalizing
the design of the catalytic surface.15,73–75 It is already widely accepted
that the optimum range to ensure the effective catalytic functional-
ity of gold is less than 5 nm.76 In many studies using XRD, the aver-
age diameter of gold particles (not less than about 3 nm) was
calculated (from typical reflections of Au at 2θ = 38.20 ° and 44.4 °
based on the broadening of the diffraction peak according to the
Scherrer equation) and used to compare different ceria-based gold
catalysts. Table 10.1, calculated by Andreeva et al.,64 shows the aver-
age size and lattice parameter for gold and compares gold catalysts
for the WGSR on ceria supports doped with Al2O3 (10 and 20 wt%)
and synthesized by CP or MA.
The main crystallite size of Au in the gold–ceria (AuCe) fresh
samples was below 2 nm but an agglomeration in the used sample
was observed. Gold particles with smaller average dimensions and
higher stability after catalytic operation were obtained using MA.
The observed systematically lower Au lattice parameter (4.0789 for
bulk Au) was explained as being due to deformation during crystal
growth at relatively low temperatures as well as to the marked lattice
mismatch of the Au and ceria phases.

b1469_Ch-10.indd 511 4/8/2013 12:37:29 PM


b1469 Catalysis by Ceria and Related Materials

512 D. Andreeva, T. Tabakova and L. Ilieva

Table 10.1 Average size and lattice parameter of gold in fresh and used catalysts.
Average size (nm) Lattice constant (Å)
Catalysts Fresh Used Fresh Used
AuCe < 2.0 (HRTEM) 6.5 — 4.073(5)
AuCeAl10CP 3.1 4.5 4.058(5) 4.055(5)
AuCeAl20CP 5.9 7.4 4.060(5) 4.065(5)
AuCeAl10MA 2.9 3.0 4.063(5) 4.064(5)
AuCeAl20MA 3.5 3.9 4.058(5) 4.047(5)

Gold with very small dimensions cannot be detected by XRD.


Ultra-small gold particles and small clusters have been evidenced by
FTIR measurements using CO as a probe molecule.13,49,66 Au parti-
cles can hardly be distinguished in TEM micrographs since the
nanocrystalline ceria and Au nanoparticles have similar contrast.41,49
Because of this low contrast between Au and CeO2 particles with
similar size, Karpenko and co-authors used an Au/CeO2 catalyst
with lower BET surface area (containing Au nanoparticles of the
same size as the higher surface area catalyst) for TEM evaluation of
the WGS reaction-induced modifications in Au particle size.25 Gold
particles less than 3 nm have been observed using annular dark field
STEM and EDS.16 The presence of very small gold particles (includ-
ing gold clusters around 1 nm), which are undetectable by XRD,
were registered using HRTEM and EDS analysis. The EDS measure-
ments focused on regions containing highly contrasting spots found
by TEM; however, as already noted, due to the difficulty of detecting
gold on nanocrystalline ceria, the particle size distribution was
not provided.13,19,50,72 A statistical analysis of particle size distribution
can be made by observations obtained with a transmission electron
microscope equipped with a Z-contrast annular detector (the so-
called high-angle annular dark field (HAADF) images). Using this
technique, it was established that some (but not very substantial)
differences in gold particle size distribution depend on the prepara-
tion method of ceria supports doped with rare earth (RE) metals51,52
or with Fe and Mn.54 Nanosized gold particles on MnOx-doped ceria,

b1469_Ch-10.indd 512 4/8/2013 12:37:29 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Gold Catalysts 513

having high thermal stability after two-step calcination (at 400 °C


and then at 550 °C), were evidenced on the basis of a particle size
distribution obtained by Z-contrast and thermal diffuse scattering
(TDS) observations (based on measurements of more than 700
particles).77
There is currently a debate about the oxidation state of Au
species, due to its importance for elucidation of the nature of the
active sites and the possible interaction of gold with ceria and
modified ceria supports. XPS has been widely used to probe the
oxidation state of gold. Fu and co-authors17,78 investigated the
oxidation state of Au in fresh and leached samples of Au/ceria and
Au/La- or Gd-modified ceria. They established that after leaching
the samples are free of metallic gold nanoparticles and only
cationic gold (Au1+ and Au3+) is present; however, their WGS activ-
ity results were the same as with high-content gold samples.
Karpenko et al.25 investigated the deactivation of an Au/CeO2
catalyst using XPS measurements after a WGS reaction with differ-
ent gas feed compositions. The contribution of cationic gold in
used samples decreased by different amounts for idealized refor-
mate, H2-rich idealized reformate, and H2-rich CO2-containing
idealized reformate. However, the authors concluded that the
decreasing number of Aun+ species cannot be the main reason for
deactivation. The presence of Au0 has been evidenced in many
studies.18,19,26 Andreeva et al. compared the WGSR over gold cata-
lysts supported on CeO2 as well as on mixed ceria supports, pre-
pared by CP or MA, containing different amounts of Al2O3 (10 or
20 wt%)64 or rare earths (10 wt%).51,79 They calculated the binding
energy (BE) for Au 4f7/2 XPS peaks of fresh samples having most
of the gold in the metallic state. Another amount with higher BE,
related to positively charged small gold particles, was also observed.
After catalytic operation, the portion of positively charged gold
fully disappeared and only metallic gold was registered in the used
samples (see Table 10.2).
Tabakova and co-workers investigated the state of gold loaded on
ceria by DP or MDP.72 The samples were pre-reduced in situ prior to
measuring the XPS spectra and gold in a metallic state was observed.

b1469_Ch-10.indd 513 4/8/2013 12:37:29 PM


b1469 Catalysis by Ceria and Related Materials

514 D. Andreeva, T. Tabakova and L. Ilieva

Table 10.2 XPS data (Au 4f7/2 region) of fresh and used (after WGSR) gold
catalysts.
Fresh samples Used samples
Catalyst Peak position (eV) At % Peak position (eV) At %
AuCe 84.70 0.56 84.15 0.45
85.90 0.12
AuCeAl10CP 84.14 0.31 84.04 0.35
85.69 0.07
AuCeAl10MA 84.24 0.29 84.27 0.34
85.77 0.06
AuCeAl20CP 84.20 0.21 83.92 0.24
86.00 0.03
AuCeAl20MA 84.26 0.27 84.14 0.27
85.95 0.03
AuCeSm10CP 84.45 0.44 84.07 0.47
85.61 0.29
AuCeSm10MA 84.25 0.11 84.17 0.24
85.91 0.15
AuCeYbMA 85.14 0.38 84.16 0.30
86.95 0.13

Both electronegative (Auδ−) and electropositive (Auδ+) oxidation


states in addition to Au0 were distinguished when studying gold on a
CeO2–MnOX mixed support prepared by mechanochemistry.77
These results are an example of the enhanced possibility of electron
density transfer at the periphery of small gold particles on different
zones of a mixed ceria support. Burch28 pointed out that XPS results
can differ from observations obtained under real reaction condi-
tions. The in situ characterization of gold–ceria catalysts by X-ray
absorption near-edge structure (XANES) and extended X-ray
absorption fine structure (EXAFS) were proposed for investigating
catalysts during a reaction.16,27 Jacobs et al.19,80 have used in situ X-ray
absorption spectroscopy to clarify the state of gold and ceria under
reducing conditions. They reported that in fresh catalysts Au is

b1469_Ch-10.indd 514 4/8/2013 12:37:29 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Gold Catalysts 515

oxidized but it is fully reduced by hydrogen at 100 °C for 5% Au on


ceria, and at 150 °C for 1% Au on ceria. The reduction of the sup-
port occurs only after Au0 formation. Using a combined EXAFS and
density functional theory (DFT) study of an Au–CeZrO4 catalyst,
Tibiletti et al. showed that zerovalent gold clusters containing about
50 atoms exist under the conditions of the WGSR.81 In another
almost simultaneous EXAFS study, it was established that at tempera-
tures above 250 °C, the WGSR occurs at the gold–ceria interface, the
active phase involving metallic gold nanoparticles smaller than
2 nm.82 Starting from a fully cationic gold state, where only Au–O
coordination is observed by EXAFS in the as-prepared samples,
Deng et al.29 studied low-content gold–ceria catalysts (< 1% Au)
under different WGS conditions. Reduction of the oxidized gold
species was evidenced, depending on the reaction gas composition.
When the reduction potential of the gas stream was lower, more
oxidized gold was detected in the used catalyst. The results support
the idea that the oxidation state of small gold particles in the border
between ceria and modified ceria supports can change depending
on the environment.

10.3.2 State of ceria and modified ceria supports


10.3.2.1 Structural features
The BET surface areas of the supports have been reported in almost
all studies of gold–ceria-based catalysts. The influence of the sup-
port surface area on the activity and deactivation behavior of Au/
CeO2 catalysts in the WGS reaction using a dilute water gas mixture
has been investigated. It has been established that Au mass normal-
ized activity increases linearly with increasing support surface area
up to 78 m2 g−1, and then remains about constant with further
increases in surface area. The stability increases with surface area up
to the same value of 78 m2 g−1, and then decreases.24 This study24 is
consistent with work by the same group22 and by Andreeva6 where
an optimum ratio of gold loading and support surface area was
shown for Au/CeO2 catalysts with variable Au loading and constant

b1469_Ch-10.indd 515 4/8/2013 12:37:30 PM


b1469 Catalysis by Ceria and Related Materials

516 D. Andreeva, T. Tabakova and L. Ilieva

ceria surface. More attention is paid in the literature to ceria disper-


sion as the nano-scaled particle size of the support is closely related
to high catalytic performance. The average crystallite size of CeO2 is
usually calculated using the broadening of the CeO2 (111) reflec-
tion in XRD patterns from the Scherrer equation. The cell param-
eters are more usually calculated from an average of several planes,
such as (111), (200), (220), (113), and (311). Fu et al. ascribed the
very high activity of nanostructured Au/CeO2 catalysts in the low-
temperature WGS reaction to the enhanced reducibility of ceria
particles with high dispersion.14,15 Carrettin et al. established that
a nanocrystalline ceria support (with a mean ceria particle size of
3–4 nm) enhanced the PROX activity of Au by two orders of magni-
tude compared to Au deposited on a regular cerium oxide prepared
by precipitation.39
Numerous studies have focused on gold catalysts on ceria sup-
ports doped by other metals since gold dispersion is usually higher
using modified ceria supports. The more defective ceria structure
is expected to cause higher catalytic activity. XRD reflections of
doped ceria shift to lower degrees with the incorporation of cations
with a larger ionic radius than Ce4+ (such as La3+ or Sm3+) and to
higher degrees with the incorporation of cations with a smaller
ionic radius (such as Zn2+).50 Many studies have confirmed that the
presence of doping metals, especially in solid solution structures,
inhibits the sintering of the CeO2 particles.17,20,56,64 An increase in
the lattice constant was observed in gold–ceria compared to gold-
free ceria samples,17,64 which was explained by the substitution of
gold ions in the ceria lattice and the presence of Ce3+ (having a
larger ionic radius compared to Ce4+) as well as the formation of
oxygen vacancies due to ceria modification by gold. The existence
of oxygen vacancies would also expand the ceria lattice according
to the literature.83 For doped ceria, the presence of only a solid
solution-like phase or the existence of a second separate phase of
the dopant depends on the nature and the concentration of the
dopant as well as on the synthesis method of the mixed support.
Estimation of the ceria lattice constant is informative for ceria
modification depending on the dopant’s ionic radius and the

b1469_Ch-10.indd 516 4/8/2013 12:37:30 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Gold Catalysts 517

preparation mode. Fu et al. established that La- and Gd-doping (10


at% La and 10 or 30 at% Gd, urea gelation/co-precipitation
method of preparation) increases the lattice constant due to the
expansion of the ceria lattice by the substitution of Ce4+ with the
larger La3+ and Gd3+ ions and that the corresponding particle size
of gold-doped ceria is considerably smaller than that of gold-free
ceria.17 The expansion of the lattice with an increase of the
La-dopant concentration (up to 12 at%, CP method of prepara-
tion) was reported by Zhang et al.47 It was also established that on
increasing the La loading, the ceria particle size decreased gradu-
ally. By studying gold on ceria doped by modifiers with the same
atomic ratio (M/(M + Ce) = 0.05 where M = Sm, Zn, or La, CP
preparation method), Avgouropoulos et al.49 found that Sm3+ and
La3+ cations were fully incorporated into the ceria lattice leading to
the formation of a solid solution, whereas most of the zinc oxide
species were highly dispersed on the ceria surface. Similar results
concerning the ineffective introduction of Zn as a dopant were
observed by Laguna and co-authors56 when comparing ceria oxides
doped with Zn, Zr, and Fe (10 mol%, prepared by the pseudo-sol–
gel method). Supports containing Zr and Fe had a lower cell
parameter compared to pure ceria; however, the sample containing
Zn had the same cell parameter and the presence of the ZnO phase
was evidenced. Tabakova et al.58 calculated lower values of the CeO2
lattice constants for Ce–Fe mixed oxide materials in comparison
with pure ceria. The decrease of the lattice parameter originated
from contraction of the cell due to the substitution of Ce4+ (0.97 Å)
by Fe3+, which has lower ionic radius (0.64 Å) and implied the for-
mation of a cubic CeO2-like solid solution. The data show that the
calculated unit cell parameter (0.5340 nm) and ceria particle size
(4.1 nm) were a minimum for a support containing 25 wt% of
Fe2O3. The cell dimension has the same value (0.5343 nm) inde-
pendent of the increase of the Fe2O3 content from 25 to 50 wt%.
This means that the addition of a higher amount of Fe2O3 did not
lead to an increased amount of solid solution; however, a new
hematite phase appeared. The results are consistent with those
reported by Laguna et al. that 25% iron content is the solubility

b1469_Ch-10.indd 517 4/8/2013 12:37:30 PM


b1469 Catalysis by Ceria and Related Materials

518 D. Andreeva, T. Tabakova and L. Ilieva

limit of iron in a ceria matrix.57 The XRD spectra of gold catalysts


supported on ceria doped with FeOx, MnOx, and CoOx (10 wt%),
synthesized by CP or by MA, differ.54 For these samples, CeO2 dif-
fraction lines typical of the fluorite-type structure of ceria were
registered; however, double phases were observed with Fe- and
Co-containing catalysts prepared by MA. The extent of the scatter-
ing of the lattice constant for the XRD peaks was correlated with
the strain caused by insertion of the dopant. The values of the lat-
tice constant showed a larger contraction of the ceria lattice for CP
catalysts containing Fe and Co compared to the corresponding MA
samples (Table 10.3). This is in agreement with the effect of the
preparation method. Co-precipitation leads to deeper insertion of
dopant into the ceria structure, while predominantly surface modi-
fication is caused by MA. The same value of the a parameter was
obtained with CP and MA catalysts containing Mn. The peak-broad-
ening strain b was also similar for both preparation methods. This
could be explained because Mn can easily enter the octahedral,
ceria-type oxygen arrangement, as in an MnO2 crystal pyrolusite of
the same stoichiometry as CeO2 (the most probable oxidation state
of the dopant ions was estimated to be Mn4+ using the TPR results).54
A possible way of elucidating the defectivity of the ceria surface
is by estimation of the oxidation state of the cerium (Ce4+ or Ce3+)
by means of XPS. Before catalytic work, the relative intensity of the

Table 10.3 Lattice constant (a) and strain


parameter (b) evaluated by XRD.
Sample a (Å) b
AuCe 5.411(4) 0.010
AuCeFeCP 5.394(7) 0.011
AuCeFeMA 5.404(7) 0.007
AuCeMnCP 5.394(3) 0.006
AuCeMnMA 5.394(3) 0.009
AuCeCoCP 5.378(3) 0.015
AuCeCoMA 5.408(3) 0.012

b1469_Ch-10.indd 518 4/8/2013 12:37:30 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Gold Catalysts 519

Ce3+ related peaks, normalized to the total Ce(3d) intensity,


depends on the pretreatment procedure. As expected, oxidative
pretreatment leads to a decrease, while reductive pretreatment
leads to an increase of the Ce3+ concentration.22,27,72 Jacobs et al.80
found the Ce3+ amount increased to 19.2% in a 1 wt% Au/CeO2
catalyst during reductive treatment in H2/He at 200 °C using the in
situ XANES technique. After reductive pretreatment of Au/CeO2 at
200 °C, Karpenko et al.24 reported that the oxidation state of ceria
depends on the particle size — with larger ceria particles the num-
ber of surface Ce3+ ions decreased. On the basis of results obtained
for the characterization (including XPS results for Ce3+ concentra-
tion in fresh and used samples) and the WGS activity of gold–ceria
and gold–ceria–alumina catalysts prepared by different methods
and containing different Al amounts, Andreeva et al. proposed a
model, discussed below, of the active sites for WGS, including the
participation of surface Ce3+ in close contact with small gold
nanoparticles and oxygen vacancies.64 However, Karpenko and
co-authors obtained results that do not support a general correla-
tion between WGS activity and cationic Aun+ species or Ce3+ species/
oxygen vacancies.25
Most researchers agree that defects (i.e. oxygen deficiencies) on
the surface of ceria play an important role in the WGS and PROX
mechanism. Many studies have been reported in the literature recom-
mending Raman spectroscopy as a method for characterizing defects
in ceria structure. However, there are fewer papers that describe the
use of Raman bands to obtain reliable information about defects in
ceria supports with gold catalysts for WGS and PROX reactions. The
most intense Raman line at about 460 cm−1 is ascribed to the Raman
active F2g mode (the single allowed Raman mode in metal oxides with
a fluoride structure) and corresponds to the oxygen symmetric
breathing vibration around Ce4+ ions.84 The presence of oxygen
vacancies and the existence of reduced Ce3+ species are usually con-
nected to a low intense Raman band located at 580–600 cm−1 (the
detection and relative intensity of the band are affected by the laser
power and frequency, the particle size distribution and homogeneity
as well as the absorption properties of the solid).85–87 The full width at

b1469_Ch-10.indd 519 4/8/2013 12:37:30 PM


b1469 Catalysis by Ceria and Related Materials

520 D. Andreeva, T. Tabakova and L. Ilieva

half maximum (FWHM) of the main ceria band depends on the ceria
particle size as well as on defects in the ceria lattice, mainly related to
the presence of dopant ions or oxygen vacancies.88,89 The values of the
FWHM reported by Andreeva et al. studying gold catalysts supported
on ceria doped with different amounts of Al2O3 (10 or 20 wt%)20,64
show that the addition of alumina, either by CP or MA, causes widen-
ing of the main ceria line. The FWHM values are larger for CP com-
pared to MA catalysts. The higher broadening with CP samples may
be due to the higher dispersion of the CeO2 particles and to the more
defective structure caused by the co-precipitation. Solids with higher
amounts of alumina, synthesized by the same preparation mode
(CP or MA), demonstrate higher broadening. Raman spectroscopy
data for RE-containing gold catalysts (10 wt%)51,52 differ from those
obtained with alumina as dopant. Figure 10.1 shows Raman spectra of
the initial ceria supports modified with La, Sm, Gd, and Y using
different methods.
The weak band detected at about 600 cm−1 is evidence for the
defective ceria structure induced by both CP and MA. However, in
the case of initial supports prepared by MA, the FWHM data for the
main ceria band suggested a larger number of defects. These
defects could be mainly oxygen vacancies because the rare-earth
(3+) ions, applied as dopants, are with ionic radii that suppose a
vacancies formation in order to attain the neutrality. The calcu-
lated FWHM values for MA samples are in contrast to the expecta-
tion of higher defectivity caused by co-precipitation as with
Al-containing ceria supports. The average size of ceria crystallites
was quite close for both preparation modes, so any dependence of
the FWHM on the dispersion has to be excluded. The proposed
explanation was that most probably the oxygen vacancies were adja-
cent to the RE3+ dopant and the ceria structure seems to be better
ordered than ceria modified with alumina. The changes in the posi-
tion of the main ceria band are informative for the insertion of the
dopants and the position shift is visible only for the RE-containing
supports prepared by CP (Fig. 10.1). In this case the dependence
of the frequency on the dopant ionic radius was linear. No such
correlation was found when using MA.51 The results are in

b1469_Ch-10.indd 520 4/8/2013 12:37:30 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Gold Catalysts 521

Figure 10.1 Raman spectra of initial ceria supports doped with REs: (a) prepared
by CP and (b) prepared by MA. Reprinted from Ilieva et al.52 with permission from
Elsevier.

agreement with the suggestion of deep ceria modification by CP


and predominantly surface modification by MA. Raman data col-
lected for the characterization of the ceria supports of gold cata-
lysts after addition of dopants (Sm3+, La3+, and Zn2+)50 also show a
shift of the main ceria line toward lower frequencies and its broad-
ening compared to pure ceria. The small band detected at about
580–590 cm−1 has been attributed to the presence of defects (vacan-
cies or impurities) in the ceria structure. Laguna and co-authors56

b1469_Ch-10.indd 521 4/8/2013 12:37:30 PM


b1469 Catalysis by Ceria and Related Materials

522 D. Andreeva, T. Tabakova and L. Ilieva

compared a series of ceria oxides doped with 10 mol% of Zr, Zn,


and Fe (using the pseudo-sol–gel method). In this study the num-
ber of oxygen vacancies was evaluated using Raman data by measur-
ing the ratio between the areas of bands associated with oxygen
vacancies and the main F2g band (a higher ratio meaning a higher
number of oxygen vacancies). A similar number of vacancies as
observed for pure ceria was obtained with the Zn-doped support,
which is in good accordance with the low Zn–Ce interaction and
tendency of ZnO segregation evidenced by XRD, as already dis-
cussed. For the iron-doped ceria supports, the shift towards lower
frequencies and the broadening of the F2g band on increasing iron
content (10, 25, and 50 mol%, pseudo-sol–gel method of prepara-
tion) was registered.57 However, the band assigned to the presence
of oxygen vacancies in the structure (598–600 cm−1) disappeared
for the doped systems (Fig. 10.2).

Figure 10.2 Raman spectra of the oxides. Reprinted from Laguna et al.57 with
permission from Elsevier.

b1469_Ch-10.indd 522 4/8/2013 12:37:30 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Gold Catalysts 523

The Fe-modified sample had the lowest number of oxygen


vacancies, nevertheless the changes in the position and the width
of the main ceria band were the strongest and they clearly demon-
strate solid solution formation. Similar behavior has been reported
by Bao and co-workers for Ce–Fe mixed oxides prepared by co-
precipitation.90 They established that the band for the oxygen
vacancies disappears when the Fe molar ratio reaches 0.4 and
above in the Ce–Fe solid. The authors reported that small doping
amounts of Fe3+ facilitate the formation of oxygen vacancies,
whereas large doping amounts annihilate them. The results56,57,91
show that the modification of the ceria structure is strongly
dependent on the nature of the dopant (especially the radius of
the doping ion). After doping ceria with Me3+ ions, the crystal
lattice can compensate for the excess negative charge with three
mechanisms: vacancy compensation, cerium interstitial compensa-
tion, and dopant interstitial compensation.11,91 For small cations
with radius < 0.8 Å, vacancy compensation is accompanied by some
degree of dopant interstitial compensation; for Me3+ dopants with
larger radii, vacancy compensation mainly applies.11 When two
Ce4+ ions are substituted by two trivalent Me3+ ion with larger radii
(RE metal ions, for example) then via the vacancy compensation
mechanism, one oxygen vacancy is created to balance the charge.
Dopant interstitial compensation prevails when the dopant ion
radius is smaller than Ce4+ as in the case of Fe3+. Three Ce4+ ions
are substituted by three Fe3+ ions and neutrality is accomplished by
an additional Fe3+ ion located at the interstitial sites of the cubic
CeO2 structure. Vacancy compensation promotes the generation
of oxygen vacancies while dopant interstitial compensation elimi-
nates them. The processes depend on the dopant concentration.
Bao et al.90 established that within the range of low Fe3+ concentra-
tion, doping uses vacancy compensation and thus the concentra-
tion of oxygen vacancies increases with an increasing amount
of doped Fe3+. However, when the concentration of doped Fe3+
exceeds a critical value, the mechanism switches to dopant intersti-
tial compensation, in which increasing the amount of doped Fe3+
annihilates oxygen vacancies.57,58,91

b1469_Ch-10.indd 523 4/8/2013 12:37:30 PM


b1469 Catalysis by Ceria and Related Materials

524 D. Andreeva, T. Tabakova and L. Ilieva

10.3.2.2 Redox behavior of gold–ceria and gold-doped ceria catalysts


The redox activity of ceria-based materials is well known and the
effect of gold as well as of the doping ions on the Ce4+ ↔ Ce3+ tran-
sition has been widely investigated by means of H2-TPR measure-
ments. The TPR profile of pure ceria consists of two peaks — the
first TPR peak with a maximum at about 500 °C is due to the reduc-
tion of surface ceria layers, and the next peak with Tmax in the
interval 700–800 °C is due to bulk reduction.92 The key finding for
Au/CeO2 catalysts for the low-temperature WGS reaction is that
the surface oxygen of ceria is substantially weakened by the pres-
ence of gold nanoparticles. This was reported by the research
groups of Andreeva — at the First Gold Conference in Cape Town,
South Africa, 2001, and later published in Catalysis Today12 — and
Flytzani-Stephanopoulos, who observed that the degree of weak-
ness of the oxygen depends on the preparation method, calcina-
tion temperature, and gold loading.14 The authors pointed out that
the amount of surface oxygen in Au/CeO2, available for reduction,
is controlled by the crystal size of ceria, which was in agreement
with the already published observation of Trovarelli et al.11 Figure 10.3
illustrates the shift of temperature for ceria surface reduction by
several hundred degrees to lower temperatures with a maximum at
about 100 °C.12
The TPR peaks are complex — hydrogen consumption (HC)
could be due to two processes: the reduction of the oxygen species
coordinated around the small gold particles or to the reduction of
ceria surface layers. The visible differences in peak shape and posi-
tion depend on the gold content in the samples. The improved
reduction of surface ceria can be explained on one hand by
enhanced oxygen mobility at the border between ceria and nano-
gold particles and on other by the evidenced possibility of small gold
particles to activate hydrogen by dissociation.93 The reduction of
bulk oxygen in CeO2 was found to remain unchanged by the pres-
ence of gold.12,14 These general observations were confirmed by
TPR results in many other investigations.15,20,41,45,55,64 The surface
reduction processes monitored by TPR in combination with XANES

b1469_Ch-10.indd 524 4/8/2013 12:37:30 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Gold Catalysts 525

Figure 10.3 Low-temperature TPR profiles of samples with different gold loading.
Reprinted from Andreeva et al.12 with permission from Elsevier.

and in situ diffuse reflectance IR Fourier-transform spectroscopy


(DRIFTS) allowed Jacobs and co-authors19 to conclude that the
reduction of ceria is facilitated by the formation of Au0. The authors
proposed a mechanism for ceria surface reduction including the
generation of bridging OH groups (formed either by water dissocia-
tive adsorption at the sites of oxygen deficiency or directly by spill-
over of dissociated H2 from the gold to the ceria surface). Each pair
of these OH groups leads to a change of the oxidation state of two
surface ions from Ce4+ to Ce3+.19
The effect of ceria doping on the reduction behavior of the
initial supports and corresponding gold catalysts depends on the
dopants’ nature and on the applied preparation method. Using
irreducible dopants (Al and some RE metals), the changes

b1469_Ch-10.indd 525 4/8/2013 12:37:30 PM


b1469 Catalysis by Ceria and Related Materials

526 D. Andreeva, T. Tabakova and L. Ilieva

550
535

Detector responce
510 855
150
Detector responce

CeAl20 x 64
CeAl10 x 64
CeO2 x 64

100 200 300 400 500 600 700 800 9001000


114 Temperature, °C
x 128 120 x 32 AuCeAl20
AuCe

x 128 x 32
AuCeAl10
x 128 x 32

0 100 200 300 400 500 600 700 800 900 1000
Temperature, °C

Figure 10.4 TPR profiles of gold catalysts on ceria and ceria–alumina supports
prepared by CP. Inset: TPR of the corresponding initial supports. Reprinted from
Andreeva et al.20 with permission from Elsevier.

in reducibility compared to undoped ceria can be related to the


modification of the ceria structure. Figure 10.4 shows the TPR
results reported by Andreeva et al.20 with ceria/alumina supports (10
or 20 wt% Al2O3) prepared by CP and with gold catalysts.
It can clearly be seen (Table 10.4) that the addition of alumina
leads to higher H2 consumption (HC), e.g. enhanced oxygen mobil-
ity in the ceria supports (20 wt% Al2O3 has a much higher HC). The
results show that oxygen mobility is significantly increased by the
insertion of Al3+ by CP. Another reason for the easier reduction
could be the lowering of the average size of the ceria particles
caused by this preparation method. The reduction of the ceria sur-
face layers, which does not affect the fluorite structure of the grain,
is considered to be limited to 17%94 or 20%.95 Bearing in mind this
supposition, the TPR peaks are due not only to the reduction of
ceria surface layers, but also to reduction in the bulk (especially for
the support with 20 wt% alumina — the calculated degree of reduc-
tion was 43.3%). For samples containing gold, the first TPR peak

b1469_Ch-10.indd 526 4/8/2013 12:37:30 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Gold Catalysts 527

Table 10.4 HC corresponding to the first


peak in the TPR profiles of the samples.
Sample HC (mmol g−1)
Ce 0.49
CeAl10CP 0.72
CeAl20CP 1.01
CeAl10MA 0.30
CeAl20MA 0.35
AuCe 0.46
AuCeAl10CP 0.53
AuCeAl20CP 0.57
AuCeAl10MA 0.35
AuCeAl20MA 0.32

was recorded at very low temperatures. However, the HC values for


AuCeAl10CP and AuCeAl20CP are lower than for the corresponding
support. Oxygen mobility depends on the temperature and within
the range for surface ceria reduction it cannot be as high as 400–
600 °C for the initial supports. The higher oxygen storage capacity
(OSC) due to the presence of alumina is also seen for gold
catalysts.
Table 10.4 also shows the values of HC obtained for the sup-
ports and gold catalysts prepared by MA.64 The HC is lower than
for the corresponding samples prepared by CP. The gold catalysts
prepared by MA exhibited better WGS activity than those
prepared by CP (an explanation is given in Section 10.4.2.1).
Figure 10.5 shows and Table 10.5 summarizes TPR results after
RT and high-temperature (HT) re-oxidation (after the first TPR
peak).64
For CP gold catalysts re-oxidation in the deeper ceria layers can-
not be completely achieved even during HT re-oxidation (compare
with the TPR data listed in Table 10.4). With MA, the OSC remains
practically unchanged during the redox cycle into the temperature
range of interest for WGSR.64

b1469_Ch-10.indd 527 4/8/2013 12:37:31 PM


b1469 Catalysis by Ceria and Related Materials

528 D. Andreeva, T. Tabakova and L. Ilieva

Figure 10.5 TPR profiles of gold catalysts on ceria doped with alumina (10 or
20 wt%) after re-oxidation at RT and HT: (a) samples prepared by CP and (b) sam-
ples prepared by MA. Reprinted from Andreeva et al.20 with permission from Elsevier.

Table 10.5 HC corresponding to the first TPR peak after re-oxida-


tion (RO) at RT or at HT.
HC (RO at RT) HC (RO at HT)
Catalyst (mmol g−1) (mmol g−1)
AuCe 0.24 0.30
AuCeAl10CP 0.36 0.36
AuCeAl20CP 0.32 0.40
AuCeAl10MA 0.22 0.32
AuCeAl20MA 0.22 0.33

b1469_Ch-10.indd 528 4/8/2013 12:37:31 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Gold Catalysts 529

Andreeva et al.51 observed higher WGS activity also over gold


catalysts on ceria modified with RE (RE=Y, La, Sm, Gd, or Yb) pre-
pared by MA compared to the corresponding CP-prepared gold
catalysts. No distinct correlation between reducibility and WGS activ-
ity was established in this case. In contrast, for the same two series of
catalysts, the gold catalysts prepared by CP performed better in
PROX than those synthesized by MA.52 At the relatively low tempera-
tures of the PROX reaction, surface oxygen vacancies play a major
role and their generation is enhanced by the lower temperature
reducibility of the CP samples (see Fig. 10.6).
Due to the predominant surface modification of ceria by MA,
the presence of larger amounts of dopant on the surface causes dis-
turbance between the gold and the ceria. The well-resolved double
TPR peaks of the MA samples (Fig. 10.6) support this assumption,
showing that some of the modified ceria is located far from the gold
particles. These results could explain why MA gold catalysts are less
active in PROX than their CP analogues. A single and narrow

103

118

108 149
Detector responce
Detector responce

117
x2 AuCeYCP 155 AuCeGdMA

x2 94 AuCeGdCP 140
108 109 AuCeSmMA
x2 AuCeSmCP

AuCeLaMA
x2 AuCeLaCP

0 50 100 150 200 250 0 50 100 150 200 250


Temperature (oC) Temperature (oC)
(a) (b)

Figure 10.6 Low-temperature TPR patterns of gold catalysts on ceria doped with
RE: (a) prepared by CP and (b) prepared by MA. Reprinted from Ilieva et al.52 with
permission from Elsevier.

b1469_Ch-10.indd 529 4/8/2013 12:37:31 PM


b1469 Catalysis by Ceria and Related Materials

530 D. Andreeva, T. Tabakova and L. Ilieva

TPR peak was registered with the AuCeYCP sample, which could
mean that the surface is homogeneous and there is close contact
between the gold and the ceria. It is interesting to comment that this
catalyst exhibited the highest activity in PROX52 and the lowest activ-
ity in WGSR.51 This was reasonably explained as being due to the
redox properties of the catalysts at different temperatures. After
re-oxidation at 220 °C (the temperature which is of interest for
WGS)51 a very low value for hydrogen consumption of the AuCeYCP
sample was observed. Compared to other RE-containing catalysts, it
had the largest difference in HC between fresh and re-oxidized sam-
ples. However, after re-oxidation at 100 °C (the temperature of inter-
est for PROX) oxygen capacity almost fully recovered, being only a
little lower compared to the fresh sample.52 Wang et al.55 observed a
correlation between WGS activity and the reducibility of gold cata-
lysts supported on ceria doped with different La amounts. The
authors observed that the TPR peak area for ceria surface reduction
initially increased significantly with La doping between 10% and
25% and then dropped with further La addition. This is the same
dependence on La content as in catalytic activity in WGS.
When the dopants are also reducible, the assignment of the
reduction processes is more complicated as changes to the oxida-
tion states of the dopants’ ions have to be taken into consideration
as well. The TPR processes for gold catalysts for WGS or PROX sup-
ported on ceria doped with transition metals have been studied in
relation to the nature and concentration of the dopant as well as the
method of preparation.48,52,56,59
Fewer studies deal with CO-TPR, nevertheless this technique
could be more informative concerning the catalytic behavior for
WGS and PROX reactions. Fu and co-workers showed that in the
presence of gold, the surface reduction of ceria by CO begins below
room temperature and several surface oxygen species of different
reducibility were distinguished.78 The drop in PROX selectivity with
temperature can be explained by the results showing that the tem-
perature of H2 oxidation is higher than that of CO oxidation on
gold–ceria, as evidenced by comparison of H2- and CO-TPR.15,17
Studying gold on La-doped ceria Jain et al.45 performed CO-TPR

b1469_Ch-10.indd 530 4/8/2013 12:37:31 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Gold Catalysts 531

measurements and the results showed that the reduction tempera-


ture decreased by approximately 40 °C compared to the maximum
temperature observed during H2-TPR. The temperature for high
PROX activity was correlated with the temperature for the catalyst’s
reduction by CO.

10.4 WGS Catalytic Activity on Gold–Ceria Catalysts


As already mentioned, there a few important factors for achieving
good performance of WGS gold–ceria catalysts: preparation method,
the dispersion of the gold, and the state and nature of the supports.
The latter is of crucial importance for the preparation of not only
active but also very stable catalysts for WGS reactions. For this reason
ceria is a very good choice.

10.4.1 WGS reaction on gold–ceria


Recently many authors have discussed different aspects of the
WGS reaction over gold-based catalysts on ceria: activity, stability,
deactivation, influence of reaction conditions, pretreatment of the
catalysts, etc.9,14–19,24,25,27–29,31 Andreeva’s group was a pioneer in these
studies.4,6,12,13 In 2001 during the First Gold Conference in Cape
Town, South Africa, for the first time Andreeva reported results for
gold catalysts based on ceria for WGS reactions, published later in a
special issue of Catalysis Today.12 In this publication the effect of the
gold content, space velocity, and vapor/CO ratio on WGS activity
and stability were discussed. The catalytic activity of fresh samples
and after stability tests are summarized in Table 10.6. It can be seen
that the activity of fresh samples increases with an increase in gold
loading (1 wt%), but there are no big differences between the cata-
lytic activity of the samples containing 3 and 5 wt% of gold. The
influence of space velocity on WGS activity, measured at three differ-
ent operating temperatures, was noticeable only for the sample with
low Au content. For the other two samples (3Au/CeO2 and 5Au/
CeO2) the effect of contact time on catalytic activity was negligible,
especially at higher operating temperatures. WGS activity slightly

b1469_Ch-10.indd 531 4/8/2013 12:37:31 PM


b1469 Catalysis by Ceria and Related Materials

532 D. Andreeva, T. Tabakova and L. Ilieva

Table 10.6 WGS activity of Au/CeO2 catalyst expressed as degree of CO conver-


sion at 225 °C.12
Catalysta CO conversion at 225 °C (%)
Fresh samples After 3 weeks of operation
1Au/CeO2 58.0 70.0
3Au/CeO2 68.0 82.0
5Au/CeO2 75.0 78.0
a
Note: The number before Au indicates the gold loading (wt%) of the catalysts.

increases as a function of the water vapor/CO ratio and even the


sample 3Au/CeO2 with a lower gold loading has higher CO conver-
sion in comparison with the sample with higher gold content (5Au/
CeO2). Table 10.6 also shows data obtained after different treatment
and a long period of operation. The results reveal stable catalytic
behavior. Notice that the activity of the 3Au/CeO2 sample is higher
than the 5Au/CeO2 sample. A high oxidation capacity for catalysts
based on ceria was also established, which could explain the high
and stable activity of gold–ceria catalysts.
Flytzani-Stephanopoulos and co-workers studied Au–CeO2 cata-
lysts with different gold loading.14,29 They also prepared catalysts with
a very low gold content after leaching, free of metallic gold.17 In both
cases they established high and stable low-temperature activity in
WGS reactions. They stressed the role of high oxygen capacity due
to the ceria support. The effect of nanosized ceria was also reported.
The influence of the surface area of the support on the activity
and stability of gold–ceria catalysts was studied by Behm and
co-workers.24 A wide range of surface areas of the ceria support
(24–284 m2 g−1) was investigated. It was found that gold particle size
is practically independent of the catalyst surface area. Sandoval et al.
also stressed the role of the support for good performance of a gold-
based catalyst.9 The effects of temperature and atmosphere pretreat-
ment on the activity and deactivation of gold–ceria catalysts was
reported by Karpenko and co-workers.25,27,31 The higher WGS activity
of gold–ceria catalysts is due to the strong modification of ceria

b1469_Ch-10.indd 532 4/8/2013 12:37:31 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Gold Catalysts 533

in the presence of gold. The ease of electron transfer between nano-


sized gold and ceria particles via oxygen vacancies is the main rea-
son for the higher activity of these catalysts at low temperatures.12,13
It was also observed that reactive oxygen species are controlled by
the crystal size of the ceria.15 The high activity and stability of gold
catalysts supported on ceria in WGS reactions make these catalysts
promising for practical applications.

10.4.2 WGS reaction over gold based on doped ceria


10.4.2.1 Gold WGS catalysts on ceria–alumina
Investigations by Andreeva’s group clearly show that the ceria sup-
port plays an active role in the WGS activity of gold catalysts.6,12,13
This finding motivated further studies on ceria modified by alu-
mina. The addition of alumina to ceria supports has two important
aims: to increase the number of oxygen vacancies in the ceria
structure and to improve the stability of both gold and ceria parti-
cles during catalytic operation.20,21,64 Mixed oxide supports were
prepared by three different techniques: CP, washcoating (W), and
MA.20,21 Gold catalysts with different alumina loading were tested,
but here we report only results for samples with lower Al content
(10 wt%), which exhibited higher WGS activity. A comparison of
the catalytic activity of the samples prepared by different methods
are presented in Fig. 10.7.
The results showed that the applied method of preparation of
modified ceria supports causes significant differences in WGS activ-
ity. The addition of alumina by CP results in more oxygen vacancies
than in ceria but they are formed deeper inside the ceria lattice. In
this case water vapor is not able to re-oxidize the catalyst surface dur-
ing catalytic operation.20 The preparation of a mixed ceria–alumina
support by MA leads to an increase in the number of oxygen vacan-
cies compared to gold–ceria samples. However, these vacancies are
located to a greater extent on the catalyst surface. Re-oxidation of
the catalyst surface by water vapor during WGS reactions in this case
should be easier. This could explain the higher WGS activity of gold

b1469_Ch-10.indd 533 4/8/2013 12:37:31 PM


b1469 Catalysis by Ceria and Related Materials

534 D. Andreeva, T. Tabakova and L. Ilieva

100
AuCeAlMA
90 AuCe

80
AuCeAlW
CO conversion, %

70
60
50 AuCeAlCP

40
30
20
10
0
Catalysts

Figure 10.7 WGS activity of gold catalysts supported on ceria and ceria modified
by alumina, prepared by different methods.

catalysts supported on ceria–alumina (MA) compared to the other


gold supported catalysts. The effect of the pretreatment procedure
has been shown to have less effect than the preparation method, but
higher activity was observed after pretreatment in air.20,21
Galletti et al. reported the results of WGS activity over Au sup-
ported on CeO2–Al2O333 under realistic stream conditions. The sam-
ples reached a maximum of 37% CO conversion at 330 °C, very far
from the equilibrium curve. Note, however, that above 300 °C, the
WGS activity of unpromoted ceria is significant.96
For MA samples after re-oxidation at RT and at HT (the working
temperature for WGS reactions), the maximum in the TPR peak
shifted to lower temperatures, especially after HT re-oxidation.21
The lowest temperature (80 °C) was registered for AuCeAlMA, the
catalyst exhibiting the highest WGS activity. Higher hydrogen con-
sumption (after RT as well as after HT re-oxidation) was observed
when mechanochemical activation of the mixed support was used.
These results correlate with the higher WGS activity of MA samples.
Results obtained by XPS measurements are also in agreement with
this conclusion. The addition of alumina as a structural modifier of
ceria-supported gold catalysts prevents the agglomeration of gold
and ceria nanoparticles.

b1469_Ch-10.indd 534 4/8/2013 12:37:31 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Gold Catalysts 535

10.4.2.2 Gold WGS catalysts on ceria doped by RE metals


Recently, gold WGS catalysts based on ceria doped with different
rare earth metals have been studied. In the laboratory of Prof.
Flytzani-Stephanopoulos, gold catalysts on gadolinium promoted
ceria have been studied. After removing the metallic gold, no posi-
tive effect on WGS activity was found.37 Zhang et al. studied gold
catalysts based on ceria doped by La.47 La doping can improve WGS
activity and stability by modifying the ceria structure. Recently, the
effect of La-doped ceria as a support of gold for WGS reactions was
reported by Wang et al.55 A series of catalysts with different gold
loading under different conditions were studied. A strong increase
in reactivity for an Au/LaxCe1-xO2–0.5x catalyst, prepared at an opti-
mum pH = 8.5 was found.
Mixed CeO2–Me2O3 (Me=La, Sm, Gd, or Yb) supports were
used successfully as supports of gold catalysts for WGS reactions.51,79
The mixed oxide supports were prepared by CP or MA. Results of
catalytic activity tests for WGS reactions of fresh catalysts and after
pretreatment in air are given in Figs. 10.8 and 10.9.51 It can be seen
that the MA samples are more active than those prepared by CP but
the differences are not as large as for Al-modified ceria catalysts.20,21,64

100
AuCeYbMA
90 AuCeSmMA
80 AuCeSmCP
CO conversion, %

70 AuCe AuCeGdMA
AuCeLaMA AuCeGdCP
60
50 AuCeLaCP
40
30
20
10
0
Catalysts

Figure 10.8 WGS activity of gold catalysts based on doped ceria (as prepared)
at 240 °C.

b1469_Ch-10.indd 535 4/8/2013 12:37:31 PM


b1469 Catalysis by Ceria and Related Materials

536 D. Andreeva, T. Tabakova and L. Ilieva

100
90 AuCeYbMA
AuCeSmCP
AuCe AuCeSmMA
80
AuCeLaMA AuCeGdMA
CO conversion, %

70 AuCeGdCP
60 AuCeLaCP

50
40
30
20
10
0
Catalysts

Figure 10.9 WGS activity of gold catalysts based on doped ceria activated in air
at 240 °C.

The activation of the catalysts in air leads to very interesting results:


the WGS activities of the CP-prepared samples increase, even the
sample doped by Sm (AuCeSmCP) showed higher activity than the
corresponding MA one. This experimental fact was unexpected.
A supposition is that the vacancies are localized around the RE
dopant and the ceria structure seems to be better ordered than for
Al-doped samples, which results in easier re-oxidation of the cata-
lysts during catalytic operation. Obviously, the nature of the dopants
and the way they are introduced into the ceria play a more substan-
tial role. Zhao and Gorte,97,98 who studied RE-doped ceria as catalysts
for n-butane oxidation, found that Sm-doped catalysts show the larg-
est increase in ionic conductivity and the largest decrease in activity
for n-butane oxidation from various RE dopants. The authors sup-
posed that it is energetically more favorable to have 3+ ions migrate
to the surface of a cubic fluorite structure.99 It is possible that the
surface could be enriched with Sm3+. For WGS gold catalysts doped
with Sm, the surface amount of Ce3+ was higher compared to all
other samples studied, as estimated by XPS.51 Therefore, most prob-
ably the higher activity of the doped catalysts is because the highest
ionic conductivity of ceria is in the presence of Sm. No distinct
correlation between reducibility and WGS activity was found.

b1469_Ch-10.indd 536 4/8/2013 12:37:31 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Gold Catalysts 537

10.4.2.3 Gold WGS catalysts on ceria doped by transition metals


Recently, it has been shown that CeO2–ZrO2 is a very effective sup-
port for noble metals. A comparative study of WGS reactions over
gold and platinum supported on ZrO2 and CeO2–ZrO2 was carried
out by Boaro et al.10 It was found that at lower temperatures gold
catalysts were more active than platinum ones. The addition of
zirconia to ceria leads to a modification of the redox and structural
characteristics of the supports and this affects the activity of these
catalysts.
The effect of loading ceria supports for gold WGS catalysts with
different transition metals (Fe, Mn, Sn) has been studied. Two series
of gold catalysts were synthesized by CP and MA. Generally, the MA
samples exhibited higher activity than the CP ones. Such behavior
has been observed for gold catalysts supported on ceria doped by Al
and RE metals.20,21,51,64 Significant differences in the WGS activity of
gold catalysts on ceria modified by transition metals synthesized by
MA and CP were established. The MA catalysts, doped by Fe and
Mn, manifested higher WGS activity than the undoped Au/ceria
sample and the sample doped by Sn.
The WGS activity order at 250 °C over catalysts doped by Fe, Mn,
and Sn and prepared by both techniques was as follows:
AuCeFeMA≈AuCeMnMA>AuCe>AuCeSnMA>AuCeFeCP>
AuCeMnCP>AuCeSnCP
A 40-h catalytic test at 250 °C over the AuCeFeMA sample
showed good stability of the catalyst. Insignificant deactivation was
observed. The degree of CO conversion was the following: 94.67%
after 8 h, 93.34% after 16 h, 92.01% after 24 h, and 88.2% after 32 h
as well as after 40 h.
It is complicated to estimate the contribution of each active
phase to the total WGS activity. Considering that MA induces surface
enrichment by the dopants, for this method most oxygen vacancies
are at the surface, which could be assumed to be a key factor for the
higher WGS activity. CP leads to a strong modification of ceria, while
MA causes modification mainly at the surface. The differences in
catalytic activity are due to the method of preparation and to the

b1469_Ch-10.indd 537 4/8/2013 12:37:31 PM


b1469 Catalysis by Ceria and Related Materials

538 D. Andreeva, T. Tabakova and L. Ilieva

nature of the dopant. Raman spectroscopy data show that ceria dop-
ing leads to the formation of defects, thus causing higher oxygen
mobility, i.e. a higher oxygen capacity than for a gold catalyst on
undoped ceria, as proven by TPR results. Oxygen capacity is affected
by the preparation technique as well as by the nature of the dopant.
Tabakova et al.59 studied the influence of the amount of iron
oxide in ceria on WGS catalytic activity.19 Mixed oxide supports were
prepared by urea gelation co-precipitation. The following order of
activity was observed:
Au/CeO 2 >Au/Ce50Fe50>>Au/Ce75Fe25>Au/Ce25Fe75>
Au/Fe2O3
Differences in gold particle size and in the concentration of sur-
face Ce3+ defect sites could explain the WGS activity order. The
lower oxygen vacancy concentration and Ce3+ amount in all gold
catalysts supported on mixed CeO2–Fe2O3 materials in comparison
to Au/CeO2 could be responsible for the observed performance.
In conclusion, catalysts prepared by MA are much more active
than the CP ones even when the dopants are not reducible and not
active in the WGS reaction (e.g. alumina and irreducible lantha-
nide oxides as dopants). Figure 10.10 compares results for WGS

100 AuCeFeMA AuCeMnMA


AuCeAlMA
90 AuCeSmCP
AuCeSmMA
AuCe
80
CO conversion, %

70
60
AuCeFeCP
50
AuCeAlCP
40
30 AuCeMnCP
20
10
0
Catalysts

Figure 10.10 WGS activity of gold catalysts based on ceria doped by Al, Sm, Fe,
and Mn, prepared by different methods (catalytic activity data from Andreeva
et al.20,21,51).

b1469_Ch-10.indd 538 4/8/2013 12:37:32 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Gold Catalysts 539

activity, obtained in the laboratory of Andreeva, over the more


active gold catalysts on unpromoted ceria and ceria promoted by
Al, Sm, Fe, and Mn, prepared by different techniques. It can be
seen that depending on the nature of the dopants and the prepara-
tion method applied, the observed WGS activity was different
under the same conditions of catalytic operation. In all cases the
dopants lead to an increase in the stability of the gold catalysts
based on ceria. In summary, these investigations are very useful
not only from a fundamental point of view but also for practical
application.

10.5 Active Sites and Mechanism of the WGS


Reaction Over Ceria-Based Gold Catalysts
The nature and structure of active sites in Au/ceria catalysts
under WGS reaction conditions have been experimentally and
theoretically studied in recent years. Different characterization
tools have been applied to provide relevant information and to
clarify where and how the WGS reaction steps occur on gold–
ceria-based catalytic systems. The first detailed FTIR study
focused on investigating the chemisorption and reactivity of the
molecules involved in the forward and reverse WGS reaction at
different temperatures and after different pretreatments.13 CO
as a probe at the temperature of liquid nitrogen (90 K) after
oxidative treatment produced a band at 2100 cm−1 due to CO
adsorbed on step sites of metallic gold. The low intensity of this
band in combination with HRTEM and EDS analysis suggested
the presence of very small gold clusters (d ≤ 1 nm) covered by
adsorbed oxygen after the calcination treatment in air. More
importantly, it was found that these nanosized gold particles
caused strong modification of the surface properties of the
ceria. The significantly increased intensity of the absorption
bands of CO on Ce4+ observed with Au/CeO2, in comparison
with the corresponding band on pure ceria, evidenced a large
number of co-ordinatively unsaturated sites near the very small
gold clusters. A broad absorption at 2060 cm−1 in the spectrum

b1469_Ch-10.indd 539 4/8/2013 12:37:32 PM


b1469 Catalysis by Ceria and Related Materials

540 D. Andreeva, T. Tabakova and L. Ilieva

of Au/CeO2 after reduction with hydrogen was assigned to CO


adsorbed on negatively charged gold clusters. The formation of
Auδ− sites was suggested to arise after interaction of hydrogen
with both oxygen adsorbed on the gold clusters and oxygen
atoms on the ceria surface in close contact with the gold, pro-
ducing water and oxygen vacancies or Ce 3+ defects. The presence
of these defects allows electron transfer from the support to the
gold clusters. Firstly, Boccuzzi et al.93 reported the formation of
negatively charged gold sites after reduction of Au/α–Fe2O3 and
Au/TiO2. However, these sites were detected only on the freshly
reduced as-prepared catalysts. They were not observed on used
catalysts, probably as a consequence of the sintering of small
clusters on the TiO2 and Fe2O3 supports. In contrast, Auδ− sites
were found after many redox procedures on the Au/CeO2 cata-
lyst. This finding experimentally confirms the ability of ceria to
maintain a high dispersion of the supported gold. Moreover,
from the results of in situ WGSR experiments up to 350 °C, a
band at 2100 cm−1 was detected after cooling to RT. Very similar
values were calculated for the integrated areas of this band and
for the band at 2060 cm−1 in the spectrum of freshly reduced Au/
CeO 2, which indicated that almost all active gold sites detected
at the beginning of WGS reaction were still available. The usual
position of the band for CO adsorbed on gold metallic particles
after the WGS reaction revealed that the gold sites exposed at
the surface of the small clusters were negatively charged on the
freshly reduced sample, while they were neutral after the reac-
tion and their surface concentrations were almost constant.
Additional evidence for the presence of oxygen vacancies or
Ce3+ defects on a reduced Au/ceria catalyst was produced after the
admission of oxygen on preadsorbed CO over a reduced sample.
An immediate depletion of the band at 2060 cm−1 occurred,
accompanied with the appearance of a new, more intense band at
2127 cm−1 and a weak band at 2343 cm−1 due to the formation of
CO2. The band at 2127 cm−1 was due to CO co-adsorbed with
oxygen species on gold step sites, close to the Ce3+, where oxygen

b1469_Ch-10.indd 540 4/8/2013 12:37:32 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Gold Catalysts 541

was activated and reacted with CO to produce CO2, according to


the scheme:
2060 cm-1 2127 cm-1

CO CO Ox-
δ
Au - Ce3+ + O2 → Au Ce4+

Based on all these considerations, the conclusion is that the


WGS reaction occurs at the boundary between the small metallic
gold particles and the ceria, where CO adsorption on the gold and
H2O dissociation on oxygen-vacancy defects in the ceria takes place.
Further FTIR measurements have shown that the preparation tech-
nique strongly influences WGS activity due to the large differences
in gold particle size and availability of active gold sites at the
surface.72 Figure 10.11 shows FTIR spectra collected after CO
adsorption on oxidized (a) and reduced (b) Au/ceria catalysts
prepared by DP and MDP (discussed in Section 10.2.2).

2153
2151 2137

2157
0.2 A 2143
0.1 A
2140

2060
2170

2148 2100 2078

2200 2150 2100 2050 2000 2200 2150 2100 2050 2000 1950
-1 -1
Wavenumbers (cm ) Wavenumbers (cm )
(a) (b)

Figure 10.11 FTIR absorbance spectra of 2.5 mbar CO adsorbed at 90 K on (a)


oxidized samples and (b) reduced Au/CeO2 samples, in the carbonylic region: DP
prepared (bold curves), MDP prepared (thin curves), and CeO2 (dotted curves).
Reprinted from Tabakova et al.72 with permission from Elsevier.

b1469_Ch-10.indd 541 4/8/2013 12:37:32 PM


b1469 Catalysis by Ceria and Related Materials

542 D. Andreeva, T. Tabakova and L. Ilieva

In the spectra for MDP-prepared catalysts only a very weak and


broad band at 2078 cm−1 was observed. The integrated area of this
band was 1.72, one order of magnitude lower than the integrated
area of the band at 2060 cm−1 in the spectrum of a DP-prepared cata-
lyst (18.31). These features are in good agreement with XPS data
showing lower Au concentration on the surface of the less active
MDP-prepared gold–ceria catalyst.
XPS has been used to analyze the oxidation state of gold in WGS
catalysts prepared by deposition of gold nanoparticles on mixed
ceria–alumina supports, synthesized by different methods.21,64 The
calculated BE of the Au 4f7/2 peak for the as-prepared samples shows
that most of the Au was in the metallic state but there was also a
small amount with higher BE, which could be due to positively
charged small gold particles. After the catalytic operation this part
fully disappeared and only metallic gold was found. The calculated
surface concentrations of Ce3+ before and after the WGS reactions
correlated well with catalytic activity order. Additionally, XRD pat-
terns of the as-prepared samples and samples used in the WGSR
showed that the lattice parameter of ceria is systematically higher
than that of the pure CeO2 phase. The reason for the observed posi-
tive deviations of the unit cell parameters of the studied samples was
assumed to be the formation of a CeOy phase with y ≈ 1.98.64 The
existence of a complex Auδ+ Vo Ce3+ in the as-prepared catalysts was
suggested.21 An optimum ratio between Ce3+, oxygen vacancies, and
positively charged nanosized gold particles in the investigated cata-
lysts was considered to be necessary for high WGS activity. Under the
reaction conditions this complex changed because of electron trans-
fer between the gold and ceria through the vacancies Auδ+ → Au0,
due to the dynamic equilibrium in the redox reaction studied. The
model shown in Fig. 10.12 has been proposed for the active sites for
the WGS reaction over Au/ceria catalysts.64
An interesting experiment was reported by Fu et al. for identify-
ing the active sites for the WGSR over Au/ceria catalysts.78 The
metallic gold particles were removed from the catalyst by cyanide
leaching and activity before and after leaching was compared. The
authors found that nonmetallic gold species strongly associated

b1469_Ch-10.indd 542 4/8/2013 12:37:32 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Gold Catalysts 543

Figure 10.12 Model for the WGS reaction on a gold–ceria catalyst. The 14-atom
Au cluster is deposited on the (001) plane of oxygen deficient CeOy (y < 2.0), con-
taining Ce4+ and Ce3+ ions. V0: oxygen vacancy; Au0: neutral Au atom; Auδ+: posi-
tively charged gold; Ow: water oxygen atom. Reprinted from Andreeva et al.64 with
permission from Springer.

with the ceria contributed to the excellent WGS activity of the gold–
ceria catalysts and they considered the metallic gold nanoparticles
to be spectator species only. Manzoli et al. used FTIR measure-
ments on the same catalysts to reveal the presence of cationic gold
species in a strong interaction with the ceria.100 The main observa-
tions were that in oxidizing conditions the CO interaction produces
absorption bands in the region typical for cationic gold clusters
(2130–2140 cm−1), whereas in reducing conditions zerovalent gold
clusters are produced, possibly negatively charged by strong interac-
tions with the reduced ceria and characterized by a CO IR band
red-shifted to below 2100 cm−1.
In a very recent study Guan et al. examined the interaction of
gold atoms with CeO2 nanocrystals having rod and cube shapes.101
Gold nanoparticles with sizes in the range 2–6 nm were observed
after deposition-precipitation. In addition to these nanoparticles,
EXAFS analysis showed that a substantial amount of the gold is
present as cations that replace Ce ions at the surface of the ceria
nanorods and could not be removed by cyanide leaching. After
reduction these isolated Au atoms formed finely dispersed Au clus-
ters with a high activity in the WGS reaction. The higher activity of

b1469_Ch-10.indd 543 4/8/2013 12:37:32 PM


b1469 Catalysis by Ceria and Related Materials

544 D. Andreeva, T. Tabakova and L. Ilieva

the leached catalyst in comparison with the unleached catalyst


was maintained up to a reasonably high temperature (15% CO
conversion at 300 °C), after which, probably, more extensive reduc-
tion accompanied with the formation of larger gold nanoparticles
occurred.
Kim and Thompson applied the same leaching procedure and
calculated significantly higher rates for unleached, nanocrystalline
gold-containing catalysts than for the leached catalyst.102 On the
basis of turnover frequencies (five times higher for the unleached
catalysts than for the leached catalyst) the importance of nanocrys-
talline gold for the WGS reaction was suggested. Karpenko et al. also
studied the role of cationic and nonionic Au species in the WGSR
on Au/CeO2 catalysts by comparing the performance of a cyanide-
leached catalyst (according to the procedure of Fu et al.78) with that
of unleached catalysts.26 The authors observed significant changes
of the gold species during catalyst treatments after the leaching pro-
cess (drying, calcination, heating up to reaction temperature in
inert gas, particularly during the WGS reaction). By using in situ
spectroscopic and structure-sensitive techniques (DRIFTS, TEM,
XPS) the formation of Au0 species was evidenced, including both
small aggregates and metallic nanoparticles, and their participation
in the WGS reaction was suggested. In situ EXAFS and XANES
experiments, carried out by Tibiletti et al. on highly active Au/
CeZrO4 catalysts, showed that under the WGS reaction only Au–Au
distances of 2.76 Å and 3.80 Å existed, indicating that the gold pre-
sent is in the Au0 state.81 The combination of these results with DFT
calculations supported a model based on the assumption that small
metallic gold particles in close contact with the oxide support pro-
vided the active sites for the WGSR.
In situ characterization with X-ray absorption spectroscopy was
used to determine the chemical state of the gold during the WGSR
over an Au/ceria catalyst. The data indicated that cationic Auδ+ spe-
cies present in fresh AuOx/CeO2 catalysts do not exist under reac-
tion conditions and they could not be the active sites.82 The authors
concluded that the rate-determining steps for the WGS at tempera-
tures above 250 °C occur at the gold–ceria interface, with the active

b1469_Ch-10.indd 544 4/8/2013 12:37:33 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Gold Catalysts 545

phase involving small gold clusters (< 2 nm) and O vacancies. In a


recent review of gold catalysts Rodriquez summarized that the active
phase of an Au/ceria catalyst under WGS reaction conditions con-
sists of metallic gold nanoparticles on partially reduced ceria.103
A very recent FTIR study combined with HRTEM measure-
ments gave further insights and information about the nature of
gold active species over Au/CeO2 and Au/Zn–CeO2.104 The cata-
lytic behavior in the WGSR and PROX reactions of these catalysts
was evaluated and compared. Au/CeO2 demonstrated superior
WGS activity, whilst Au/Zn–CeO2 was more active in the PROX
reaction. Spectroscopic data collected after different pretreat-
ments, in different atmospheres (H2, D2), as well as after reduction
at temperatures relevant for the WGS and PROX reactions evi-
denced that: (i) gold able to interact with CO is deposited as highly
dispersed clusters (about 1 nm) on the surface of Au/CeO2, while
small gold particles are predominantly available on the surface of
Au/Zn–CeO2; (ii) the concentration of Ce3+ defective sites, respec-
tively oxygen vacancies, is higher on the surface of Au/CeO2 than
on Au/Zn–CeO2. The modification of ceria by Zn probably causes
the formation of Au–Cex alloy clusters over a reasonable WGSR
temperature region. These alloy clusters could be responsible for
depletion of oxygen vacancies and diminished ability in water acti-
vation. Using FTIR experiments with 18O2 an enhanced supply of
active lattice oxygen during the CO + O2 interaction was observed
on Zn-doped ceria reduced at 100 °C. The findings of this study
provided additional evidence that gold clusters and Ce3+ defects are
the active sites for the WGSR reaction, while the step sites of gold
particles are important both for CO and oxygen activation and play
a decisive role in the PROX reaction.
Theoretical studies of Au on CeO2 (111), which typically con-
sider the interaction of the surface with individual Au atoms, are
based on periodic DFT. In the first density functional calculations to
consider the adsorption of gold on ceria Liu et al.105 suggested that
at the atop site of a perfect CeO2 (111) terrace Au acquires some
positive charge leading to a reduced Ce3+ cation. On the basis of this
assumption the authors demonstrated preferential Au adsorption

b1469_Ch-10.indd 545 4/8/2013 12:37:33 PM


b1469 Catalysis by Ceria and Related Materials

546 D. Andreeva, T. Tabakova and L. Ilieva

and nucleation at O vacancies and showed that only Auδ+ can adsorb
CO sufficiently strongly for subsequent catalysis. The combination
of several such Auδ+ ions in the vicinity of an O vacancy formed ultra-
small Au clusters that were thought to be the active sites for the WGS
reaction.
Results from Hernández et al. show that Au adatoms at the CeO2
(111) surface can adopt Au0, Au+, or Au− electronic configurations
depending on the adsorption site.106 Charge transfer from the gold
atom to the Ce cation occurs for gold adsorption on top of the
surface oxygen and in a bridge position between two surface oxygen
atoms, when adsorption is strongest. Adsorption at an oxygen
vacancy site is very strong and involves the formation of an Au−
anion. In discussing the advantages and drawbacks of the different
approaches that can be used to model the properties of ceria sur-
faces, Branda et al. demonstrated that the theoretical answer to the
problem of the oxidation of atomic gold on the atop site of CeO2
(111) is indeed critically affected by the choice of computational
approach.107 The conclusion of their calculations was that a very
small energy difference between neutral and cationic Au adsorbed
at the surface strongly suggests the statistical distribution of the two
species at the most stable adsorption sites of the CeO2 (111) surface
as a function of temperature.
Summarizing recent advances, it seems that all results support
the earlier conclusion of Burch28 that the active sites for the WGSR
over Au/ceria catalysts are the metallic gold in the form of very
small clusters in intimate contact with the ceria support through
interaction with surface vacancies.
There is still no consensus concerning the mechanism for the
WGSR over Au/ceria catalysts. Two reaction mechanisms have
mainly been considered in the literature — the so-called redox
mechanism and the associative mechanism, which needs further
discussion for the type of intermediate: bicarbonate, formate, or
carboxyl.28,108 According to the redox mechanism, the WGS reaction
proceeds via consecutive steps of reduction and oxidation at the
catalyst’s surface.14 These steps include: CO adsorption on the
metal surface and interaction with lattice oxygen in the ceria to

b1469_Ch-10.indd 546 4/8/2013 12:37:33 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Gold Catalysts 547

form CO2 and water dissociation on the reduced support sites with
release of H2.
The Ce-mediated redox mechanism is:

CO + Au → CO—Au
CO—Au + 2CeO2 → CO2 + Ce2O3 + Au
H2O + Ce2O3 → 2CeO2 + H2

Later, a modification was suggested to favor a mechanism relying on


nonmetallic Au species in close association with ceria as the active
sites for the WGS reaction.78
The associative mechanism involves interaction between a CO
molecule and surface hydroxyl groups and the formation of
adsorbed surface intermediates (formate, carbonate, bicarbonate,
carboxyl), which further decompose to reaction products, CO2 and
H2. The production of IR bands typical of formate during the
WGSR over Au/CeO2,13 as well as over gold catalysts supported on
ceria doped by Sm and Al,53 suggested that formate species act as
reaction intermediates. Jacobs et al. conducted numerous studies
and their findings supported a mechanism involving surface for-
mate intermediates, in agreement with previous IR work by Shido
and Iwasawa.109 According to this mechanism the main reaction
intermediates are formates generated by the reaction of CO with
bridging OH groups associated with the Ce3+ defects. The decompo-
sition of formates to form CO2 proceeds via surface carbonate spe-
cies. The role ascribed to the gold is not only to catalyze the surface
reduction of ceria to generate the active OH groups, but also to
accelerate the decomposition of the surface formate intermediates,
the proposed rate-limiting step of the reaction mechanism.19
Combined transient in situ DRIFTS and temperature-programmed
desorption data also provided evidence that formate species are
reaction intermediates.22,23 The decomposition of these formates
was found to be the rate-limiting step, and included a dynamic equi-
librium between formates adsorbed on the ceria support and for-
mate species located on active sites for formate formation and

b1469_Ch-10.indd 547 4/8/2013 12:37:33 PM


b1469 Catalysis by Ceria and Related Materials

548 D. Andreeva, T. Tabakova and L. Ilieva

decomposition. Also, on the basis of a CO and water pulse experi-


ment, Haruta and co-workers supported the WGS mechanism oper-
ating via formate intermediates.16
However, the carbonate mechanism is favored by some authors.
Using a combination of DRIFTS and steady-state isotopic transient
kinetic analysis (SSITKA) Meunier et al. assessed the reactivity of the
species formed at the surface of an Au/Ce(La)O2 catalyst during
the WGS reaction.110 The analysis revealed that surface formates are
not important factors in the WGS reaction mechanism. Their role
is ascribed to “minor reaction intermediates” but not “spectators”
because they nevertheless participate in the formation of reaction
product.
Very recently Bond stressed several difficulties with the formate
mechanism and suggested the carboxyl group (−COOH) as the pre-
ferred alternative.108 For the mechanism of decomposition of the
carboxyl or formate group, Bond proposed a simpler and more
economical reaction cycle where the production of hydrogen mol-
ecules originates from the same water molecule, in contrast to the
more complex set in which both hydrogen atoms come from two
different molecules. DFT calculations also supposed carboxyl to be
the likely intermediate.105
Chen et al. used DFT to examine the redox and formate mecha-
nisms for the water gas shift reaction on Au/CeO2.111 Their analysis
identified the main problem in both mechanisms was cleavage of
the OH bond to reproduce surface oxygen atoms on the ceria sur-
face. Four possible pathways for regeneration of surface oxygen
atoms (three in the redox and one in the formate mechanism) were
studied and barriers higher than 1 eV were calculated, which means
that OH cannot easily dissociate at low temperatures.
A universal WGS mechanism was proposed by Burch.28 The
majority opinion is that the dominant mechanism depends on the
reaction conditions, in particular temperature and H2O/CO2 ratio.
The redox mechanism would be expected to prevail at higher tem-
peratures. At low temperatures and depending on the concentra-
tion of water in the reaction mixture, carbonate or formate
decomposition steps would be more important.

b1469_Ch-10.indd 548 4/8/2013 12:37:33 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Gold Catalysts 549

10.6 PROX Activity Over Gold–Ceria Catalysts


Catalysts for PROX have to be highly active at the operating tem-
perature of the fuel cells (80–100 °C). To avoid a loss of energy,
PROX catalysts also need to be highly selective, minimizing unwanted
oxidation of hydrogen. Several studies confirmed that the rate of
CO oxidation over supported Au catalysts exceeds that of H2 oxida-
tion.112,113 The higher selectivity in PROX at low temperatures using
Au-based catalysts in comparison with Pt-based catalysts was explained
by Kandoi et al.114 using DFT calculations for the Au (111) face. They
showed that a lower energy barrier for CO oxidation than for H2
oxidation exists. In real conditions CO2 and moisture are also pre-
sent as side products in the reformate stream. PROX catalysts for
practical application have to be capable of removing trace amounts
of CO from an excess of humid H2 containing significant quantities
of CO2. The usual concentration of CO2 in a real reaction stream is
about 20% and the H2O amount is about 10%. The nature of the
support affects catalyst activity and stability in the presence of CO2;
in general acidic supports were found to be more resistant to deac-
tivation than basic ones.115 In the summary below for PROX results
over gold catalysts supported on ceria and modified ceria, the gas
feed composition used will be mentioned.
Recently, Scirè and co-authors61 obtained data supporting the
hypothesis that the PROX reaction (1% CO, 1% O2, balance in H2)
on 1B metal/ceria catalysts occurs through a Mars–van Krevelen
mechanism, which involves the active lattice oxygen species of ceria
reacting with CO adsorbed and activated on the 1B metal particles
or at the metal/support interface. This agrees well with the very high
PROX activity observed in this study of gold catalysts at low tempera-
tures in agreement with the known ability of nano-gold to activate
CO.116,117 Comparing the catalytic performance in PROX of Au/ceria
and CuO/ceria catalysts Avgouropoulos et al.42 established a strong
effect for a synthesis procedure (DP producing highly dispersed and
more easily reduced gold or copper oxide species and thus produc-
ing more active catalysts compared to MDP). Au/ceria catalysts were
remarkably more active at temperatures lower than 120 °C, while

b1469_Ch-10.indd 549 4/8/2013 12:37:33 PM


b1469 Catalysis by Ceria and Related Materials

550 D. Andreeva, T. Tabakova and L. Ilieva

CuO/ceria catalysts were more selective. The presence of CO2 and


H2O (1 vol% CO, 1.25 vol% O2, 50 vol% H2, 0–15 vol% CO2, 0–10
vol% H2O, and He as balance) caused a significant decrease in cata-
lytic activity mainly in the case of gold. Studying Au and Pt catalysts
on a wide number of supports Ko et al.36 also observed that the gold
catalysts showed the highest CO conversion at low temperatures but
the selectivity decreased noticeably with increasing H2 concentration
in the feed. Schubert et al.34 studied the effect of the metal oxide sup-
port by comparing different gold catalysts. They established that Au/
CeO2 represented the best compromise for PROX activity, selectivity,
and long-term stability under the applied conditions of simulated
methanol reformate. Promising results in PROX over gold–ceria
catalysts have been reported by many other authors.35,37–41 Panzera
et al.35 obtained CO conversion at 120 °C close to 100% and CO2
selectivity of about 40% in the temperature range 80–120 °C (the
composition of the gas mixture was 48 cc.min−1 H2, 1 cc.min−1 CO,
1.5 cc.min−1 O2 in He). The authors discussed identification of the
active species and the role played by the support. They accepted the
synergistic mechanism for CO oxidation proposed by Bond and
Thompson,115,118 where CO oxidation occurs with high reaction rates
if CO, adsorbed on a gold particle, interacts with oxygen adsorbed
on a highly reducible metal oxide support. For this reason the nature
of the gold–oxide support interface is very important, in agreement
with Haruta.119 However, Panzera and co-authors35 pointed out that
the ability of metal oxides, such as CeO2, to provide reactive oxygen
reduces the dependence of catalytic activity on metal particle size
and makes the interface characteristics crucial to obtaining high
catalytic performance. The same authors also observed that the pres-
ence of CO2 in the stream negatively affected both CO conversion
and selectivity. Considering that the contribution of the reverse water
gas shift reaction (CO2 + H2 ↔ CO + H2O) is negligible under the
experimental conditions, the observed effects were ascribed to the
formation of carbonate or carboxylate species in agreement with
Bond and Thompson.115 Studying the role of the method of gold
deposition Luengnaruemitchai et al.38 reported that the PROX activ-
ity (1% CO, 0.5–2% O2, 2% CO2, 2.6% H2O, 40% H2 in He) of an Au

b1469_Ch-10.indd 550 4/8/2013 12:37:33 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Gold Catalysts 551

catalyst depends very strongly on the preparation mode (comparing


impregnation, co-precipitation, and the sol-gel method of Au load-
ing); an Au/CeO2 catalyst prepared by CP exhibited the highest
activity because it had the smallest gold particles. Under the humidi-
fied conditions, water lowered CO conversion over this catalyst at
temperatures below 100 °C (explained by the strong adsorption of
water at the active sites). However, at high temperatures H2O was
slightly favorable, providing a hydroxyl group, which is necessary for
the reaction to take place by the mechanism proposed by Haruta.120
Increasing the CO2 concentration from 2% to 20% reduced the
activity but the selectivity was not significantly affected.38 In contrast
to the observations by other authors, Deng et al.37 reported that the
stability of low-content gold–ceria catalysts for the PROX reaction in
a realistic fuel gas mixture (1% CO, 0.5% O2, 50% H2, 10% H2O,
15% CO2 in He) was excellent without any drop in activity or selectiv-
ity in a cyclic operation up to 150 °C. The importance of the disper-
sion of the ceria support was evidenced by Carrettin et al.39 who
showed that it is possible to achieve an increase of two orders of
magnitude of the catalytic activity in PROX (under ideal conditions
of a dry feed without CO2) over gold on ceria prepared as discrete,
well-defined nanoparticles compared to a regular cerium oxide sup-
port. The effect of morphology of the ceria nanoparticles on PROX
(under ideal conditions) is demonstrated by the activity order
obtained: Au/ceria rods > Au/ceria polyhedra >> Au/ceria cubes,
which evidences the strong shape and crystal-plane effects of
nanoscale ceria.43 The best performance was for a gold catalyst on
CeO2/nanorods in comparison to Au/polyhedra and Au/cubes,
because the nanorods have the most defective structure and the
highest reducibility. Arena et al.41 found that not only the effect of
dispersion but also the amount of residual chlorine, depending on
the preparation method, controls the strength of the gold–ceria
interaction, which results in different reactivity. The removal of
residual chlorine by washing in alkali solutions enhanced the reduc-
ibility of the active phase, thus improving the functionality of the
Au/CeO2 system in both total and preferential CO oxidation
reactions.

b1469_Ch-10.indd 551 4/8/2013 12:37:33 PM


b1469 Catalysis by Ceria and Related Materials

552 D. Andreeva, T. Tabakova and L. Ilieva

Many researchers used doped ceria supports for gold catalysts in


order to improve performance in the PROX reaction using a defec-
tive fluorite structure with increased oxygen mobility. The effect of
ceria modification with different RE metals has been discussed. The
differences in PROX activity of Au/CeLaOx catalysts (1% CO, 1% O2,
50% H2, and the balance Ar) were ascribed to differences in the gold
loading, Ce/La ratio, support crystallinity, and chloride content.
A correlation between the temperature of reduction (CO-TPR meas-
urements) and the temperature of significant activity was observed.45
Zn- and Sm-doped Au/ceria catalysts (CP preparation, Me/(Me +
Ce) = 0.05) were found to be more active than Au/ceria, whereas
La-addition had the opposite effect.49 Catalyst doping improved
resistance toward deactivation by CO2 especially for CO2 and H2O
co-addition (1% CO, 1.25% O2, 50% H2 in He; addition of 15% CO2
and 10% H2O). FTIR spectroscopic analysis performed with the
same gold catalysts50 indicated the presence of Au clusters (in agree-
ment with HRTEM data, which revealed high numbers of very small
gold clusters) on the surface of the Au/ceria as well as a higher con-
centration of Ce3+ defective sites on the surface of Au/CeO2 than on
Au/Sm- and Zn-doped ceria. The results led to the conclusion that
gold clusters and Ce3+ ions are not the active sites in the PROX reac-
tion. The step sites of small metallic gold particles are the active sites
for both CO and oxygen activation and so have a decisive role in the
PROX reaction. FTIR experiments evidenced enhanced reactivity
when PROX was performed in the presence of water at 90 K and
neither formates nor bicarbonates were observed during PROX with
or without water.50 Ilieva et al.52 compared gold catalysts supported on
ceria modified by rare earths (RE=La, Sm, Gd, and Y, 10 wt% RE2O3).
It was established that catalysts on mixed oxides prepared by CP were
systematically more active than the corresponding samples on MA
prepared supports (1% CO + 1% O2 + 70% H2 in He). A gold catalyst
on Y-modified ceria prepared by CP exhibited the highest catalytic
activity and selectivity (Fig. 10.13) as well as high stability.
No substantial differences in size distribution and the average
size of the nano-gold particles in the studied catalysts were observed.
The main reason for the differences in PROX activity was

b1469_Ch-10.indd 552 4/8/2013 12:37:33 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Gold Catalysts 553

Figure 10.13 Comparison between degree of CO conversion (filled symbols) and


selectivity toward CO2 (empty symbols) over gold catalysts on ceria and doped ceria
supports prepared by (a) CP and (b) MA. Reprinted from Ilieva et al.52 with permis-
sion from Elsevier.

considered to be due to the role of the ceria supports with the


assumption of oxygen activation through chemisorption on surface
anion vacancies in the vicinity of the boundary with the gold parti-
cles, e.g. near a CO molecule adsorbed on the gold. The eventual
explanation for the lower activity as being due to the predominant
surface modification of ceria by MA as well as the best performance
of the AuCeYCP catalyst was discussed in relation to the redox
behavior of the catalysts in Section 10.3.2.2.
Reducible dopants, which are catalytically active and selective in
the PROX reaction, have also been chosen to modify ceria supports.
A synergistic effect for conversion and selectivity (1.33% CO, 1.33%
O2, 65.33% H2, and He for balance) has been observed for Au/
MnO2–CeO2 catalysts with various Mn/Ce atomic ratios (using the
incipient-wetness impregnation method to prepare the mixed
supports) for Au/CeO2 and Au/MnO2.46 The high activity of Au/
MnO2–CeO2 catalysts is due to the coexistence of metallic and non-
metallic gold species within the nano-gold particle (< 5 nm) as well
as to the minor presence of Ce3+ surface species.46 Similarly, much
higher activity in PROX (1.0% CO + 1.0% O2 + 50.0% H2, balanced
with nitrogen) has been established for a gold catalyst on

b1469_Ch-10.indd 553 4/8/2013 12:37:33 PM


b1469 Catalysis by Ceria and Related Materials

554 D. Andreeva, T. Tabakova and L. Ilieva

Au/CeO2–Co3O4 with an appropriate Ce/Co atomic ratio (sol–gel


preparation method) compared to Au/Co3O4 and Au/CeO2.48
Laguna et al.57 established that iron-modified ceria supports contain-
ing various molar percentages of Fe (0%, 10%, 25%, and 50%,
prepared by thermal decomposition of the metal propionates)
exhibited higher PROX activity (2% CO, 1% O2, 50% H2, and N2 as
balance) than bare CeO2. The synergy between Ce and Fe was a
maximum for a gold catalyst with a support of 10% Fe. This catalyst
showed higher CO conversion than Au/CeO2 but lower selectivity
(the interaction between Au, Fe, and Ce created more active sites for
the activation of H2) at temperatures below 140 °C. There was insig-
nificant lowering of CO conversion due to the presence of CO2 and
H2O in the gas stream (10% of both CO2 and H2O). Ilieva et al.54 also
observed very high and stable CO conversion close to 100% in the
interval from room temperature up to 150 °C and stable selectivity of
about 40% over a gold catalyst on a mechanochemically prepared
ceria–FeOx (10 wt% Fe2O3) support using a dry CO/H2 mixture;
however, a significant fall in activity was caused by the addition of
CO2.77 A comparison between gold catalysts on ceria supports modi-
fied with transition metals (10 wt% MeOx, Me=Fe, Mn, or Co)54
showed better catalytic performance in PROX (under ideal condi-
tions) when prepared by MA compared to CP.54 The differences in
catalytic behavior were not due to the gold because for both prepara-
tion modes a high dispersion of gold particles with an insignificant
difference in particle size distribution was evidenced (HRTEM and
HAADF observations). The reason is because the support structure
and properties depend on the nature of the dopant and on the
preparation method. For MA, besides the gold on modified ceria,
gold on separate FeOx and MnOx phases could also contribute to the
high PROX activity and selectivity. In agreement and very recently,
while studying the reduction behavior of gold on Fe-modified ceria
using TPR and Mossbauer measurements, Ilieva et al.121 established
that the existing separate hematite phase in a catalyst prepared by
MA could participate in redox processes at relatively very low tem-
peratures (hematite to magnetite transformation below 180 °C).
PROX activity and in particular the selectivity of a gold catalyst on

b1469_Ch-10.indd 554 4/8/2013 12:37:33 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Gold Catalysts 555

MA-prepared ceria–CoOx was lower compared to a support pre-


pared by CP. Only with Co as a dopant was methane formation
detected: at temperatures ≥ 150 °C for the MA catalyst and at tem-
peratures ≥ 200 °C with a CP catalyst. This could be due to the struc-
ture of the support and the easy reduction of the separate Co3O4
phase by the MA catalyst, leading to the formation of metallic cobalt
as low as about 150 °C.122 The effect of the composition of the sup-
port on the catalytic performance in PROX was studied by Tabakova
et al.58 by varying the Ce/(Ce+Fe) ratio (co-precipitation method
of preparation), which caused a different ceria structure and defec-
tivity (ceria modification by Fe via the predominant dopant intersti-
tial mechanism was discussed in Section 10.3.2.1) as well as a
different dispersion of the gold. A support with composition 50 wt%
CeO2–50 wt% Fe2O3 was the most promising. The results of detailed
characterization by different methods showed that this composition
appeared beneficial for nucleation and the growth of highly dis-
persed gold particles (1–1.8 nm) as well as for the activation and
mobility of oxygen. The presence of Fe2O3 in the supports improves
resistance to deactivation under realistic PROX conditions. Au/
Ce50Fe50 (see Fig. 10.14) was found to be the most resistant to deac-
tivation by CO2 and water (15% CO2 and 10% H2O addition).
There is consensus of opinion that the presence of small gold
nanoparticles is a prerequisite for high catalytic activity in PROX.
The required oxidation state of gold is still under debate. The pre-
vailing opinion is that CO oxidation includes CO activation by
adsorption onto Au0. High gold dispersion is needed because
the smooth surface of metallic gold does not adsorb CO, which is
adsorbed only on steps, edges, and corner sites. The structural fea-
tures of ceria supports, depending on the preparation method and
nature and amount of the dopant, affect the growth of gold particles
as well as the ability to supply reactive oxygen species. For the PROX
reaction the oxidation of adsorbed or activated H2 is a competitive
reaction, due to the PROX selectivity of the system. Along with a
discussion of the structure–reactivity relation, another important
point is the catalysts’ resistance to deactivation by CO2 and H2O in
the PROX feed. The effect of the forward WGS and reverse WGS

b1469_Ch-10.indd 555 4/8/2013 12:37:33 PM


b1469 Catalysis by Ceria and Related Materials

556 D. Andreeva, T. Tabakova and L. Ilieva

Figure 10.14 Activity and selectivity in PROX using a gas feed (a) without and
(b) with CO2 and water. Reprinted from Avgouropoulos et al.49 with permission
from Elsevier.

reactions on the activity and selectivity could be negligible under the


PROX conditions, which are of interest for fuel cells (temperature
range 80–120 °C). The reason for the negative effect of CO2 is
mainly due to competitive adsorption blocking surface active sites.
The results summarized above show that high activity and satisfac-
tory selectivity under ideal PROX conditions (a dry feed containing
CO in excess of H2) can be obtained over selected gold catalysts on
ceria and doped ceria supports. Future efforts have to focus on
finding appropriate modifications of ceria to give improved perfor-
mance in the presence of CO2 and water.

10.7 Concluding Remarks


The results presented in this chapter on the preparation, characteri-
zation, and WGS and PROX activities of gold catalysts based on

b1469_Ch-10.indd 556 4/8/2013 12:37:33 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Gold Catalysts 557

unpromoted ceria and ceria promoted by different dopants can be


summarized as follows:

— Ceria is a very good support for the preparation of gold catalysts


with good performance in WGS and PROX reactions.
— The presence of gold leads to a modification of the structural
and redox characteristics of the ceria support.
— The addition of different dopants to ceria (Al and RE and
transition metals) increases oxygen mobility, e.g. oxygen storage
capacity.
— The modification of ceria stabilizes both gold and ceria
dispersion.
— The high and stable WGS and PROX activity of gold catalysts
based on ceria make them promising for practical applications.

Acknowledgments
The authors gratefully acknowledge the financial support of COST
Action D 36/003 and CM 0903/02.

References
1. Bond, G.C., Louis, C., Thompson, D.T., Catalysis by Gold, Imperial
College Press, London, (2006).
2. Haruta, M., Kobayashi, T., Sano, H., Yamada, N., Chem. Lett., 16 (1987)
405–408.
3. Ghenciu, A.F., Curr. Opin. Solid St. M., 6 (2002) 389–399.
4. Andreeva, D., Idakiev, V., Tabakova, T., Andreev, A., J. Catal., 158
(1996) 354–355.
5. Sakurai, H., Ueda, A., Kobayashi, T., Haruta, M., Chem. Commun.,
(1997) 271–272.
6. Andreeva, D., Gold Bull., 35 (2002) 82–88.
7. Luengaruemitchai, A., Osuwan, S., Gulabi, E., Catal. Commun., 4
(2003) 215–221.
8. Hua, J., Zheng, Q., Zheng, Y., Wei, K., Lin, X., Catal. Lett., 102 (2005)
99–108.

b1469_Ch-10.indd 557 4/8/2013 12:37:33 PM


b1469 Catalysis by Ceria and Related Materials

558 D. Andreeva, T. Tabakova and L. Ilieva

9. Sandoval, A., Gomez-Cortes, A., Zanella, R., Dıaz, G., Saniger, J.M.,
J. Mol. Cat. A: Chem., 278 (2007) 200–208.
10. Boaro, M., Vicario, M., Llorca, J., De Leitenburg, C., Dolcetti, G.,
Trovarelli, A., Appl. Catal. B, 88 (2009) 272–282.
11. Trovarelli, A., Catal. Rev. Sci. Eng., 38 (1996) 439–520.
12. Andreeva, D., Idakiev, V., Tabakova, T., Ilieva, L., Falaras, P., Bourlinos, A.,
Travlos, A., Catal. Today, 72 (2002) 51–57.
13. Tabakova, T., Boccuzzi, F., Manzoli, M., Andreeva, D., Appl. Catal. A,
252 (2003) 385–397.
14. Fu, Q., Weber, A., Flytzani-Stephanopoulos, M., Catal. Lett., 77 (2001)
87–95.
15. Fu, Q., Kudriavtseva, S., Saltsburg, H., Flytzani-Stephanopoulos, M.,
Chem. Eng. J., 93 (2003) 41–53.
16. Sakurai, H., Akita, T., Tsubota, S., Kiuchi, M., Haruta, M., Appl. Catal.
A, 291 (2005) 179–187.
17. Fu, Q., Deng, W., Saltsburg, H., Flytzani-Stephanopoulos, M., Appl.
Catal. B, 56 (2005) 57–68.
18. Kim, C.H., Thompson, L.T., J. Catal., 230 (2005) 66–74.
19. Jacobs, G., Ricote, S., Patterson, P.M., Graham, U.M., Dozier, A.,
Khalid, S., Rhodus, E., Davis, B.H., Appl. Cat. A, 292 (2005)
229–243.
20. Andreeva, D., Ivanov, I., Ilieva, L., Abrashev, M.V., Appl. Catal. A, 302
(2006) 127−132.
21. Andreeva, D., Ivanov, I., Ilieva, L., Sobczak, J.W., Avdeev, G., Tabakova, T.,
Appl. Catal. A, 333 (2007) 153–160.
22. Leppelt, R., Schumacher, B., Plzak, V., Kinne, M., Behm, R.J., J. Catal.,
244 (2006) 137–152.
23. Denkwitz, Y., Karpenko, A., Plzak, V., Leppelt, R., Schumacher, B.,
Behm, R.J., J. Catal., 246 (2007) 74–90.
24. Karpenko, A., Leppelt, R., Plzak, V., Cai, J., Chuvilin, A., Schumacher, B.,
Kaiser, U., Behm, R.J., Top. Catal., 44 (2007) 183–198.
25. Karpenko, A., Leppelt, R., Cai, J., Plzak, V., Chuvilin, A., Kaiser, U.,
Behm, R.J., J. Catal., 250 (2007) 139–150.
26. Karpenko, A., Leppelt, R., Plzak, V., Behm, R.J., J. Catal., 252 (2007)
231–242.

b1469_Ch-10.indd 558 4/8/2013 12:37:33 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Gold Catalysts 559

27. Karpenko, A., Denkwitz, Y., Plzak, V., Cai, J., Leppelt, R., Schumacher, B.,
Behm, R.J., Catal. Lett., 116 (2007) 105–115.
28. Burch, R., Phys. Chem. Chem. Phys., 8 (2006) 5483–5500.
29. Deng, W., Frenkel, A.I., Si, R., Flytzani-Stephanopoulos, M., J. Phys.
Chem. C, 112 (2008) 12834–12840.
30. Daly, H., Ni, J., Thompsett, D., Meunier, F.C., J. Catal., 254 (2008)
238–243.
31. Abd El-Moemen, A., Karpenko, A., Denkwitz, Y., Behm, R.J., J. Power
Sources, 190 (2009) 64–75.
32. Abd El-Moemen, A., Kucerova, G., Behm, R.J., Appl. Catal. B, 95
(2010) 57–70.
33. Galletti, C., Specchia, S., Saracco, G., Specchia, V., Top. Catal., 52 (2009)
688–692.
34. Schubert, M.M., Plzak, V., Garche, J., Behm, R.J., Catal. Lett., 76 (2001)
143–150.
35. Panzera, G., Modafferi, V., Candamano, S., Donato, A., Frusteri, F.,
Antonucci, P.L., J. Power Sources, 135 (2004) 177–183.
36. Ko, E.Y., Park, E.D., Seo, K.W., Lee, H.C., Lee, D., Kim, S., Catal.
Today, 116 (2006) 377–383.
37. Deng, W., De Jesus, J., Saltsburg, H., Flytzani-Stephanopoulos, M.,
Appl. Catal. A, 291 (2005) 126–135.
38. Luengnaruemitchai, A., Osuwan, S., Gulari, E., Int. J. Hydrogen Energy,
29 (2004) 429–435.
39. Carrettin, S., Concepcion, P., Corma, A., Nieto, J.M.L., Puntes, V.F.,
Angew. Chem. Int. Ed., 43 (2004) 2538–2540.
40. Goerke, O., Pfeifer, P., Schubert, K., Appl. Catal. A, 263 (2004) 11–18.
41. Arena, F., Famulari, P., Trunfio, G., Bonura, G., Appl. Catal. B, 66
(2006) 81–91.
42. Avgouropoulos, G., Papavasiliou, J., Tabakova, T., Idakiev, V.,
Ioannides, T., Chem. Eng. J., 124 (2006) 41–45.
43. Yi, G., Xu, Z., Guo, G., Tanaka, K., Yuan, Y., Chem. Phys. Lett., 479
(2009) 128–132.
44. Cargnello, M., Gentilini, C., Montini, T., Fonda, E., Mehraeen, S.,
Chi, M., Herrera-Collado, M., Browning, N.D., Polizzi, S., Pasquato, L.,
Fornasiero, P., Chem. Mater., 22 (2010) 4335–4345.

b1469_Ch-10.indd 559 4/8/2013 12:37:34 PM


b1469 Catalysis by Ceria and Related Materials

560 D. Andreeva, T. Tabakova and L. Ilieva

45. Jain, A., Zhao, X., Kjergaard, S., Stagg-Williams, S.M., Catal. Lett., 104
(2005) 191–197.
46. Chang, L.-H., Sasirekha, N., Chen, Y.-W., Wang, W.-J., Ind. Eng. Chem.
Res., 45 (2006) 4927–4935.
47. Zhang, Q., Zhan, Y., Lin, X., Zheng, Q., Catal. Lett., 115 (2007)
143–147.
48. Wang, H., Zhu, H., Qin, Z., Wang, G., Liang, F., Wang, J., Catal.
Commun., 9 (2008) 1487–1492.
49. Avgouropoulos, G., Manzoli, M., Boccuzzi, F., Tabakova, T.,
Papavasiliou, J., Ioannides, T., Idakiev, V., J. Catal., 256 (2008)
237–247.
50. Manzoli, M., Avgouropoulos, G., Tabakova, T., Papavasiliou, J.,
Ioannides, T., Boccuzzi, F., Catal. Today, 138 (2008) 239–243.
51. Andreeva, D., Ivanov, I., Ilieva, L., Abrashev, M.V., Zanella, R.,
Sobczak, J.W., Lisowski, W., Kantcheva, M., Avdeev, G., Petrov, K.,
Appl. Catal. A, 357 (2009) 159–169.
52. Ilieva, L., Pantaleo, G., Ivanov, I., Zanella, R., Venezia, A.M.,
Andreeva, D., Inter. J. Hydrogen Energy, 34 (2009) 6505–6515.
53. Andreeva, D., Kantcheva, M., Ivanov, I., Ilieva, L., Sobczak, J.W.,
Lisowski, W., Catal. Today, 158 (2010) 69–77.
54. Ilieva, L., Pantaleo, G., Ivanov, I., Maximova, A., Zanella, R., Kaszkur,
Z., Venezia, A.M., Andreeva, D., Catal. Today, 158 (2010) 44–55.
55. Wang, Y., Liang, S., Cao, A., Thompson, R.L., Veser, G., Appl. Catal. B,
99 (2010) 89–95.
56. Laguna, O.H., Romero Sarria, F., Centeno, M.A., Odriozola, J.A.,
J. Catal., 276 (2010) 360–370.
57. Laguna, O.H., Centeno, M.A., Arzamendi, G., Gandia, L.M., Romero
Sarria, F., Odriozola, J.A., Catal. Today, 157 (2010) 155–159.
58. Tabakova, T., Avgouropoulos, G., Papavasiliou, J., Manzoli, M.,
Boccuzzi, F., Tenchev, K., Vindigni, F., Ioannides, T., Appl. Catal. B,
101 (2011) 256–265.
59. Tabakova, T., Manzoli, M., Paneva, D., Boccuzzi, F., Idakiev, V., Mitov, I.,
Appl. Catal. B, 101 (2011) 266–274.
60. Adachi, G.Y., Masui, T., in Catalysis by Ceria and Related Materials, ed.
A. Trovarelli, Imperial College Press, London, (2002), pp. 51–83.

b1469_Ch-10.indd 560 4/8/2013 12:37:34 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Gold Catalysts 561

61. Scirè, S., Riccobene, P.M., Crisafulli, C., Appl. Catal. B, 101 (2010)
109–117.
62. Yuan, Z.Y., Idakiev, V., Vantomme, A., Tabakova, T., Ren, T.Z., Su, B.L.,
Catal. Today, 131 (2008) 203–210.
63. Lenite, B.A., Galletti, C., Specchia, S., Int. J. Hydrogen Energy, 36 (2011)
7750–7758.
64. Andreeva, D., Ivanov, I., Ilieva, L., Sobczak, J.W., Avdeev, G., Petrov, K.,
Top. Catal., 44 (2007) 173–182.
65. Zhang, Q., Zhan, Y., Lin, X., Zheng, Q., Catal. Lett., 115 (2007)
143–147.
66. Vindigni, F., Manzoli, M., Chiorino, A., Tabakova, T., Boccuzzi, F.,
J. Phys. Chem. B, 110 (2006) 23329–23336.
67. Manzoli, M., Vindigni, F., Chiorino, A., Tabakova, T., Idakiev, V.,
Boccuzzi, F., React. Kin. Catal. Lett., 91 (2007) 213–221.
68. Idakiev, V., Tabakova, T., Tenchev, K., Yuan, Z.Y., Ren, T.Z., Su, B.L.,
Catal. Today, 128 (2007) 223–229.
69. Idakiev, V., Tabakova, T., Tenchev, K., Yuan, Z.Y., Ren, T.Z., Vantomme, A.,
Su, B.L., J. Mat. Sci., 44 (2009) 6637–6643.
70. Si, R., Flytzani-Stephanopoulos, M., Angew. Chem. Int. Ed., 47 (2008)
2884–2887.
71. Sayle, T.X.T., Parker, S.C., Sayle, D.C., Phys. Chem. Chem. Phys., 7
(2005) 2936–2941.
72. Tabakova, T., Boccuzzi, F., Manzoli, M., Sobczak, J.W., Idakiev, V.,
Andreeva, D., Appl. Catal. B, 49 (2004) 73–81.
73. Haruta, M., Cattech, 6 (2002) 102–115.
74. Kozlov, A.I., Kozlova, A.P., Liu, H., Iwasawa, Y., Appl. Catal. A, 182
(1999) 9–28.
75. Bond, G.C., Thompson, D.T., Gold Bull., 33 (2000) 41–51.
76. Date, M., Haruta, M., J. Catal., 201 (2001) 221–224.
77. Ilieva, L., Pantaleo, G., Ivanov, I., Zanella, R., Sobczak, J.W., Lisowski, W.,
Venezia, A.M., Andreeva, D., Catal. Today, 175 (2011) 411–419.
78. Fu, Q., Saltsburg, H., Flytzani-Stephanopoulos, M., Science, 301 (2003)
935–938.
79. Andreeva, D., Ivanov, I., Sobczak, J.W., Lissowski, W., Petrova, P.,
Abrashev, M.V., Ilieva, L., Curr. Top. Catal., 7 (2008) 33–41.

b1469_Ch-10.indd 561 4/8/2013 12:37:34 PM


b1469 Catalysis by Ceria and Related Materials

562 D. Andreeva, T. Tabakova and L. Ilieva

80. Jacobs, G., Patterson, P.M., Williams, L., Chenu, E., Sparks, D.,
Thomas, G., Davis, B.H., Appl. Catal. A, 262 (2004) 177–187.
81. Tibiletti, D., Amieiro-Fonseca, A., Burch, R., Chen, Y., Fisher, J.M.,
Goguet, A., Hardacre, C., Hu, P., Thompsett, A., J. Phys. Chem. B, 109
(2005) 22553–22559.
82. Wang, X., Rodriguez, J.A., Hanson, J.C., Perez, M., Evans, J., J. Chem.
Phys., 123 (2005) 221101–221105.
83. Chen, L., Fleming, P., Morris, V., Holmes, J.D., Morris, M.A., J. Phys.
Chem. C, 114 (2010) 12909–12919.
84. McBride, J.R., Hass, K.C., Poindexter, B.D., Weber, W.H., J. Appl. Phys.,
76 (1994) 2435–2441.
85. Keramidas, V.G., White, W.B., J. Chem. Phys., 59 (1973) 1561–1562.
86. Spanier, J.E., Robinson, R.D., Zhang, F., Chan, S.-W., Herman, I.P.,
Phys. Rev. B, 64 (2001) 2454071–2454078.
87. Marban, G., Fuertes, A.B., Appl. Catal. B, 57 (2004) 43–53.
88. Graham, G.W., Weber, W.H., Peters, C.R., Usmen, R., J. Catal., 130,
(1991) 310–313.
89. Kosacki, I., Suzuki, T., Anderson, H.U., Colomban, P., Solid State Ionics,
149 (2002) 99–105.
90. Bao, H., Chen, X., Fang, J., Jiang, Z., Huang, W., Catal. Lett., 125
(2008) 160–167.
91. Trovarelli, A., ed., Catalysis by Ceria and Related Materials, Imperial
College Press, London, (2002).
92. Yao, H.C., Yao, Y.F.Y., J. Catal., 86 (1984) 254–265.
93. Boccuzzi, F., Chiorino, A., Manzoli, M., Andreeva, D., Tabakova, T.,
J. Catal., 188 (1999) 176–185.
94. Sanchez, M.G., Gazquez, J.L., J. Catal., 104 (1987) 120–135.
95. Laachir, A., Perrichon, V., Bardi, A., Lamotte, J., Catherine, E.,
Lavalley, J.C., El Faallah, J., Hilaire, L., le Normand, F., Quemere, E.,
Sauvion, G.N., Touret, O., J. Chem. Soc. Faraday Trans., 87 (1991)
1601–1609.
96. Whittington, B.I., Jiang, C.J., Trimm, D.L., Catal. Today, 26 (1995)
41–52.
97. Zhao, S., Gorte, R.J., Appl. Catal. A, 248 (2003) 9–18.
98. Zhao, S., Gorte, R.J., Appl. Catal. A, 277 (2004) 129–135.

b1469_Ch-10.indd 562 4/8/2013 12:37:34 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Gold Catalysts 563

99. Cordatos, H., Bunluesin, T., Stubenrauch, J., Vohs, J.M., Gorte, R.J.,
J. Phys. Chem., 100 (1996) 785–793.
100. Manzoli, M., Boccuzzi, F., Chiorino, A., Vindigni, F., Deng, W.,
Flytzani-Stephanopoulos, M., J. Catal., 245 (2007) 306–313.
101. Guan, Y., Ligthart, D.A.J.M., Pirgon-Galin, Ö., Pieterse, J.A.Z., van
Santen, R.A., Hensen, E.J.M., Top. Catal., 54 (2011) 424–438.
102. Kim, C.H., Thompson, L.T., J. Catal., 244 (2006) 248–250.
103. Rodriguez, J.A., Catal. Today, 160 (2011) 3–10.
104. Tabakova, T., Manzoli, M., Vindigni, F., Idakiev, V., Boccuzzi, F., J. Phys.
Chem. A, 114 (2010) 3909–3915.
105. Liu, Z.-P., Jenkins, S.J., King, D.A., Phys. Rev. Lett., 94 (2005)
196102–196104.
106. Hernández, N.C., Grau-Crespo, R., de Leeuw, N.H., Sanz, J.F., Phys.
Chem. Chem. Phys., 11 (2009) 5246–5252.
107. Branda, M.M., Castellani, N.J., Grau-Crespo, R., De Leeuw, N.H.,
Hernandez, N.C., Sanz, J.F., Neyman, K.M., Illas, F., J. Chem. Phys., 131
(2009) 094702 (1–11).
108. Bond, G.C., Gold Bull., 42 (2009) 337–342.
109. Shido, T., Iwasawa, Y., J. Catal., 141 (1993) 71–81.
110. Meunier, F.C., Reid, D., Goguet, A., Shekhtman, S., Hardacre, C.,
Burch, R., Deng, W., Flytzani-Stephanopoulos, M., J. Catal., 247
(2007) 277–287.
111. Chen, Y., Cheng, J., Hua, P., Wang, H., Surf. Sci., 602 (2008)
2828–2834.
112. Haruta, M., Tsubota, S., Kobayashi, T., Kageyama, H., Jenet, M.,
Delmon, B., J. Catal., 144 (1993) 175–192.
113. Grisel, R.J.H., Weststrate, C.J., Goossens, A., Craje, M.W.J., Van der
Kraan, A.M., Nieuwenhuys, B.E., Catal. Today, 72 (2002) 123–132.
114. Kandoi, S., Gokhale, A.A., Grabow, L.C., Dumesic, J.A., Mavrikakis, M.,
Catal. Lett., 93 (2004) 93–100.
115. Bond G.C., Thompson, D.T., Catal. Rev. Sci. Eng., 41 (1999) 319–388.
116. Haruta, M., Tsubota, S., Kobayashi, T., Kageyama, H., Genet, M.J.,
Delmon, B., J. Catal., 144 (1993) 175–192.
117. Minicò, S., Scirè, S., Crisafulli, C., Visco, A.M., Galvagno, S., Catal.
Lett., 47 (1997) 273–276.

b1469_Ch-10.indd 563 4/8/2013 12:37:34 PM


b1469 Catalysis by Ceria and Related Materials

564 D. Andreeva, T. Tabakova and L. Ilieva

118. Bond, G.C., Thompson, D.T., Gold Bull., 33 (2000) 41–51.


119. Haruta, M., Cattech 6 (2002), 102–115.
120. Haruta, M., Catal. Today, 36 (1997) 153–166.
121. Ilieva, L., Munteanu, G., Petrova, P., Tabakova, T., Velinov, N., Mitov, I.,
React. Kinet., Mech. Catal., 105 (2011) 39–52.
122. Ilieva, L., Munteanu, L., Petrova, P., Tabakova, T., Sobczak, J.W.,
Lisowski, W., Krawczyk, M., Abrashev, M.V., Andreeva, D., Nanosci.
Nanotechnol., ed. E. Balabanova, I. Dragieva, 11 (2011) 59–63.

b1469_Ch-10.indd 564 4/8/2013 12:37:34 PM


b1469 Catalysis by Ceria and Related Materials

CHAPTER 11

CERIA-BASED FORMULATIONS
FOR CATALYSTS FOR DIESEL
SOOT COMBUSTION
Eleonora Aneggi, Carla de Leitenburg
and Alessandro Trovarelli
Dipartimento di Chimica, Fisica e Ambiente, via Cotonificio 108,
Università di Udine, 33100 Udine, Italy

11.1 Introduction
11.1.1 Background
In recent years attention to air quality and the environment has
grown rapidly and the public is aware of problems originating from
urban air pollution due to traffic and vehicle emissions. Diesel and
gasoline engines are sources of the major outdoor air pollutants like
volatile organic compounds (VOCs), NOx and particulate matter.
The market penetration of diesel engines is progressively growing,
not only in Europe, where they benefit from an excellent reputa-
tion, but also in the USA where they have always been less popular.
Moreover, their market share is increasing not only in the heavy-
duty, but also in the light-duty vehicle market segments.1 The diesel
engine is the most efficient internal-combustion engine, providing
the greatest amount of energy from a specific amount of fuel. They
are also more reliable and durable than gasoline engines, and
they emit very low concentrations of hydrocarbons and carbon mon-
oxide when operating under lean regimes (air/fuel ratio higher

565

b1469_Ch-11.indd 565 4/8/2013 12:39:05 PM


b1469 Catalysis by Ceria and Related Materials

566 E. Aneggi, C. de Leitenburg and A. Trovarelli

than 20).2 The most important pollutants (excluding CO2, which is


not yet subject to specific legislative constraints) are NOx and par-
ticulates, whose reduction represents a very challenging task.
Diesel particulate matter (PM) is a complex multi-phase multi-
component material, which is traditionally and conveniently divided
into three main fractions: a solid fraction mainly comprising carbon
particles and ash, a soluble organic fraction where the majority of
the hydrocarbons originating from the lube oil and fuel are concen-
trated, and sulfate particulates derived from the sulfur content of
the fuel. The sizes of particulate matter from diesel engines are
characterized by a distribution spread (Fig. 11.1): PM10 (diameter
less than 10 µm), fine particles (less than 2.5 µm), ultrafine particles
(less than 0.1 µm) and nanoparticles (less than 50 nm).2
In contrast to gaseous diesel emissions, PM does not represent a
defined chemical species. From a legislative point of view PM is
defined as any matter in the exhaust of an internal combustion
engine that can be trapped on a specified sampling filter medium at
325 K, by drawing an exhaust gas sample from the vehicle exhaust
system and diluting with a required amount of air. The specific
composition of diesel particulate matter depends upon several
Normalized concentration

Nuclei mode
d(C/Ctot)/dDp

Accumulation mode

Coarse

1 10 100 1000 10000


Diameter (nm)

Figure 11.1 Diesel particulate size distribution. Solid line: number distribution.
Broken line: mass distribution. Adapted from Majewski and Khair.2

b1469_Ch-11.indd 566 4/8/2013 12:39:05 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Formulations for Catalysts for Diesel Soot Combustion 567

factors: engine type, age, driving conditions, speed, fuel, fuel and
lube oil additives, exhaust temperatures and sampling methods.3,4
Diesel particulate emissions are suspected to induce health
problems such as asthma, bronchitis and most importantly lung
cancer; for this reason there is an increasing attention regarding
exhaust emissions in general and on diesel soot in particular. The
National Institute for Occupational Safety and Health (NIOSH)5
and the International Agency for Research on Cancer (IARC)6
consider diesel exhaust as a potential and probable human carcino-
gen.1,7 In addition to the potential hazard for human health, par-
ticulates are also among the factors responsible for damage to the
environment by diminishing air clarity and contributing to global
warming.
The most recent Euro V legislation and the forthcoming Euro VI
(Table 11.1), as well as analogous emission limits in other parts of
the world, have introduced more stringent NOx and particulate lim-
its for light- and heavy-duty vehicles and for passenger cars, which
impose the use of combined catalyst and trap technologies to reach
the required standard.
The reduction of NOx under a permanent oxygen-rich environ-
ment is a complex issue and cannot be accomplished with three-way
catalyst (TWC) technologies where the air/fuel (A/F) ratio is main-
tained within a narrow range around the stoichiometric point.
Additionally, the reduction of particulate matter, due to its very

Table 11.1 European emission standards for light-duty engines in g.Km−1.8


Phase Year HC + NOx NOx CO TPM
Euro I 1992 0.97 — 2.72 0.14
Euro II-IDI 1996 0.7 — 1.0 0.08
Euro II-DI 1999 0.9 — 1.0 0.10
Euro III 2000 0.56 0.50 0.64 0.05
Euro IV 2005 0.30 0.25 0.50 0.025
Euro V 2009 0.23 0.18 0.50 0.005
Euro VI 2014 0.17 0.08 0.50 0.005

b1469_Ch-11.indd 567 4/8/2013 12:39:05 PM


b1469 Catalysis by Ceria and Related Materials

568 E. Aneggi, C. de Leitenburg and A. Trovarelli

complex composition, is possible only with the use of a combination


of filter traps and a diesel oxidation catalyst carefully positioned
within the engine exhaust unit.8 The combined removal of NOx and
particulates imposes the use of a combination of selective catalytic
reduction or lean NOx adsorbers with catalytic filter traps, which are
the candidate technologies to meet the legislation limits in passen-
ger cars in the near future.9,10

11.1.2 Removal of soot in diesel engine emissions


Modifications of combustion chamber design along with improved
fuel systems and the higher attention to lube oil consumption have
made significant improvements in diesel exhaust emissions in
recent years. Nevertheless, NOx and particulate matter emissions
cannot be reduced as required by current legislation through
engine modifications only. In order to fulfil the mandatory emission
requirements, gas-after-treatment technologies have been intro-
duced in recent years. The use of particulate filters and de-NOx cata-
lysts is a promising approach, following the basic concept of
three-way catalysts for gasoline engines. Soot abatement in diesel
exhaust conditions commonly consists of a two-step technology.4,11,12
The first step is the separation of carbonaceous particles from the
gas phase by mechanical filtration using a so-called diesel particulate
filter (DPF).13–15 The second step is the burning of soot particles
inside the filter in order to avoid undesired back-pressure due to the
accumulation of particulate matter in the filter. However, the elimi-
nation of diesel soot from the filter using oxygen as an oxidant
requires temperatures above 550 °C, while diesel exhaust gas tem-
peratures most of the time are between 250 and 450 °C. To achieve
regeneration of the filter by combustion of the collected particles,
the soot must reach a minimum temperature of 600–650 °C in order
to auto-ignite and sustain combustion; alternatively to lower the
required oxidation temperature and ensure complete regeneration,
a catalytic system can be employed. Catalysts can be used in a num-
ber of ways to accelerate the combustion of the accumulated soot.
They can be added either through the fuel or deposited directly in

b1469_Ch-11.indd 568 4/8/2013 12:39:05 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Formulations for Catalysts for Diesel Soot Combustion 569

the filter channels through impregnation. In either case the contact


points between the solid catalyst and the soot particles is a key
parameter for efficient and fast oxidation. There are two different
approaches for catalytic regeneration: direct and indirect contact
catalysis. In direct contact, the catalysts must be intrinsically active,
and also be able to establish intimate physical contact with the car-
bon particles.16–18 Conversely, in indirect contact catalysis, the cata-
lysts oxidize soot without the need for intimate physical contact;
they promote the formation of mobile compounds (adsorbed O
species from oxygen, NO2 from NO), which then act as the true
oxidant for soot in a more active and efficient way compared to O2.
An example of this regeneration mode is given by the oxidation of
carbon soot with NO2,19 where Pt promotes NO to NO2 oxidation
and NO2 is then transported via the gas phase over the soot parti-
cles, oxidizing carbon while being reduced back to NO.

11.1.3 Experimental approaches


The most common laboratory methods employed to investigate the
activity of soot oxidation catalysts include temperature-programmed
oxidation/combustion (TPO/TPC) and thermogravimetric (TG)
analysis of catalyst-soot mixtures. In both cases a specific amount of
soot and catalyst are loaded into a sample holder/reactor and
heated in an oxygen-containing atmosphere at a constant heating
rate. In TPO, the profile of CO2 (CO) and O2 are generally followed
during a temperature-programmed experiment (with time or tem-
perature) and the CO2 peak-temperature (usually indicated as Tm)
is used to estimate the activity of the catalyst; additional indications
of activity include the onset temperature of CO2 formation. The loss
of weight of the sample against temperature is the key parameter
when comparing catalyst activity using thermal analysis (the activity
is generally quantified as the temperature at which 50% of the soot
is oxidized, indicated as T50). Most often these results are not used
as absolute values but are quite useful for comparing a series of cata-
lyst formulations under the same experimental conditions. When
comparing activity measured in different laboratories under various

b1469_Ch-11.indd 569 4/8/2013 12:39:05 PM


b1469 Catalysis by Ceria and Related Materials

570 E. Aneggi, C. de Leitenburg and A. Trovarelli

reaction conditions, the temperature difference between unpro-


moted soot oxidation and soot oxidation promoted by a catalyst can
be conveniently used to estimate the level of activity of the formula-
tion under investigation.
The results of soot oxidation tests performed with temperature-
programmed methods (either TPO or TG) are strongly affected by
the choice of the reaction conditions. In a nice study, Peralta et al.
investigated the effect of the main parameters on the soot oxidation
reaction under laboratory conditions.20 The most important varia-
bles that might influence final results are the composition of soot
and soot/catalyst ratio, the type of soot/catalyst contact and the
amount of oxygen in the feed mixture.
A large variation of these parameters can be found in the litera-
ture and Fig. 11.2 shows how the values are distributed within the
studies analysed in this work. The use of synthetic soot is preferred
over ‘real’ soot due to a better characterization and to a better
reproducibility of results; the specific composition of real soot
depends upon several factors such as engine type, age, driving con-
dition, speed, fuel, fuel and lube oil additives, and temperature.
Moreover real soot is easier to oxidize than synthetic soot due to the
soluble organic fraction that burns at a lower temperature, thus
affecting in a conservative manner the results obtained.
A great variability in the soot/catalyst ratio is also found in the
literature, with values ranging from a soot/catalyst ratio of 1/2 down
to 1/80, with the most common values being within the range 1/5
to 1/20. An increase in the amount of soot relative to the mass of
catalyst beyond a certain value, where no more catalyst is available
to come into contact with the soot particles, will result in a shift of
the oxidation temperatures in TPO to higher values. Catalyst/soot
ratios higher than 10 do not generally influence oxidation tempera-
tures, while for a ratio of 5:1 or lower, negative effects on the oxida-
tion rate will start to be observed.20
The study of soot oxidation on a laboratory scale faces the issue
of soot–catalyst mixture preparation. This is a relevant problem since
the reaction rate is determined by the level of interaction between
the two solids and the gas mixture.3 Contact between carbon and

b1469_Ch-11.indd 570 4/8/2013 12:39:05 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Formulations for Catalysts for Diesel Soot Combustion 571

35 40
31 35
30 28 35
30
25 30
22
Percentage (%)

Percentage (%)
25
20
20
15 16
12 15
10 10
10
5 3
2 5 3
1 1 1 2 2 1
0 0
1:2 1:4 1:5 1:9 1:10 1:20 1:50 1:80 5% 6% 8% 10% 15% 21% 30% 50% O2/NOx
Soot/catalyst ratio Amount of oxygen (%)
(a) (b)
60

50

40
Percentage (%)

30

20

10

0
N330 Printex U Printex V Elftex 125 Vulcan 6 CB (Sigma MA7 Real
(Degussa) (Degussa) (Degussa) (Cabot) (Cabot) Aldrich) (Mitsubishi
Chem.
Kind of soot Corp.)

(c)

Figure 11.2 Analysis of reaction variables used in the studies analysed in this
work. (a) Soot/catalyst ratio. (b) Amount of oxygen. (c) Kind of soot (real or
synthetic).

catalyst particles is very important in solid–solid reactions and for


this reason it has been widely discussed in recent years.21–25 Three
different kinds of contact between soot and catalyst are generally
considered in the literature:4,26–28 loose contact is achieved by gently
mixing the soot and catalyst with a spatula; tight contact results from

b1469_Ch-11.indd 571 4/8/2013 12:39:05 PM


b1469 Catalysis by Ceria and Related Materials

572 E. Aneggi, C. de Leitenburg and A. Trovarelli

mechanically mixing the soot–catalyst mixture for long time (e.g. in


mortar for 10 minutes); and in situ contact is realized by filtering
diesel soot directly from a simulated or real diesel exhaust stream.
In situ and loose contact show similar results, while samples pre-
pared with tight contact exhibit better and more reproducible per-
formance due to the higher number of contact points between the
catalyst and the soot.29 Tight contact is therefore preferred when a
comparison between different types of formulation is needed and
when rate parameters need to be measured. By forcing contact
between the soot and the catalyst, the reaction becomes independ-
ent of the number of contact points and the intrinsic activity of dif-
ferent catalysts can be evaluated.3–4
Tight contact conditions are often preferred when reproducibil-
ity of results is the main target (46% of investigations) while loose
contact (29%) is preferred when more realistic behaviour is needed.
Recently, Hensgen and Stowe,25 with the aim of finding a reliable,
homogeneous and reproducible contact mode to be used in auto-
matic testing, focused on so-called ‘wet contact’. This method con-
sists in stirring the catalyst and soot in the presence of an organic
solvent for a fixed period of time. They pointed out that the benefits
of this methodology are the possibility of fine tuning by changing
the solvent and stirring time and the absence of influence due to the
operating personnel. The use of different solvents for bare ceria
samples results in a variation of T50 from 433 to 588 °C, which makes
comparison with existing data difficult.
Feed composition and flow rates also affect oxidation profiles.
The flow rate should be chosen so that oxidation occurs under a
chemically controlled regime or with negligible mass control
effects.20 Oxygen concentration within the feed mixture generally
varies from 5% to 50%, with air most often being used; normally a
shift of the oxidation profile to lower temperatures is observed when
increasing oxygen content.

11.2 Overview of the Use of Ceria for Soot Oxidation


Over the past few years, cerium dioxide and ceria-based materials
have been intensively studied as catalysts and as structural and

b1469_Ch-11.indd 572 4/8/2013 12:39:05 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Formulations for Catalysts for Diesel Soot Combustion 573

electronic promoters of heterogeneous catalytic reactions to improve


the activity, selectivity and thermal stability of catalysts.30,31 They are
used in a large number of industrial processes and account for a
large portion of the rare-earth oxide market;32 undoubtedly their
major commercial application is in the treatment of emissions from
internal combustion engines where ceria-based materials have been
used in recent decades.33,34 The utilization of ceria in TWCs15,35 for
the treatment of exhaust gases from automobiles constitutes its eco-
nomically and technologically most important application, and has
stimulated a strong research effort; however, several other processes
(fluid catalytic cracking, ammoxidation and dehydrogenation pro-
cesses30–32,36) also benefit from the use of ceria. A large amount of
work has also been carried out, especially in recent years, on the use
of ceria and related materials in reactions connected to hydrogen
production processes (reforming, water gas shift, CO preferential
oxidation) and fuel cell technology.37
The reason for the successful use of ceria in catalysis and espe-
cially in TWC has been discussed in detail in several review
papers,32,38–42 and it is connected with its ability to undergo easy, fast,
and reversible reduction to sub-stoichiometric phases. In TWCs, it is
more important that the ceria takes up and releases oxygen follow-
ing variations in the stoichiometric composition of the feedstream.
When acting in this way, the ceria is continuously and rapidly sub-
jected to reduction and oxidation cycles with its oxygen composition
alternating between CeO2 and CeO2−x. This function is called oxy-
gen storage/release capacity or more generally it is referred to as
redox behaviour, which accounts for all the characteristics of the
reduction and oxidation processes of CeO2. The capabilities of the
redox couple Ce+4–Ce+3 are strongly enhanced if other elements,
belonging for example to the rare-earth or transition-metal families,
are introduced into the CeO2 lattice by forming solid solutions. This
is the basis for the formulation of several CeO2-containing materials
for catalysis.
The preparation of catalyst formulations for soot oxidation fol-
lows different approaches. One approach is to increase the number
of contact points between the soot particles and the catalysts by using
fuel-borne catalyst additives or molten salt catalysts, which can wet the

b1469_Ch-11.indd 573 4/8/2013 12:39:05 PM


b1469 Catalysis by Ceria and Related Materials

574 E. Aneggi, C. de Leitenburg and A. Trovarelli

soot surface and therefore decrease the oxidation temperature.28,43,44


The use of a more powerful oxidant than oxygen, like NO2, can also
efficiently decrease the soot combustion temperature.45–47 Of the sev-
eral catalyst components used for soot oxidation, the most promising
formulations are based on the addition of potassium to oxides of
transition metals (such as Cu, V, Mo, Co or Fe),3,28,44,48–50 or on systems
based on a combination of Co,K/MgO or Ba,K/CeO2,46,51,52 or on
the use of perovskites like LaCoO3, La0.9Rb0.1CoO3, KxLa1−xFeO3, La1−x
CsxCoO3, LaMnO3 and La0.9K0.1Cr0.9O3.53–55
Several supports, such as ZrO2, TiO2, Al2O3 and CeO2, have also
been used in diesel soot oxidation. Ceria alone or in combination
with other oxides is active in the oxidation of carbon particles45,46,56–73
and the mechanism is due to the high availability of surface active
oxygen and high surface reducibility.60,62,63,67 Starting from the early
studies of van Doorn et al. in 1992,74 the use of ceria in soot oxida-
tion has shown a rapid increase in the last few years as a conse-
quence of the stricter regulations for particulate emissions.
A literature search with the keywords ‘ceria/CeO2/cerium dioxide’
and ‘soot oxidation’ returned more than 170 papers, whose distribu-
tion per year is shown in Fig. 11.3.
The success of oxygen storage systems based on ceria is due to
their ability to change oxidation state during operation (i.e. CeO2 to
CeO2−x) while maintaining structural integrity, thus allowing oxygen
uptake and release to occur easily. The use of CeO2-supported met-
als can increase the benefit of bare ceria due to the establishment of
an interaction between the metal and the support, which enhances
the redox characteristics of the pure ceria.46,47,60,68,75 In addition, a
mechanism in which the redox of CeO2 contributes to the genera-
tion of superoxide species in a more efficient manner compared to
other oxides has recently been put forward to explain the higher
reactivity among a series of catalyst supports.76
Although a large number of studies have appeared in the last
few years, the role of ceria as a catalyst or promoter in soot oxidation
is still a matter of discussion. The purpose of this chapter is to sum-
marize these studies with an attempt to offer the reader a consistent
picture of the role of ceria-based formulations in soot combustion.

b1469_Ch-11.indd 574 4/8/2013 12:39:05 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Formulations for Catalysts for Diesel Soot Combustion 575

30

25

Number of publications 20

15

10

0
1992
1994
1996
1998
2000
2002
2004
2006
2008
2010
Year

Figure 11.3 Number of papers on the use of ceria in soot oxidation published
1992–2011. Information taken from data found on ISI Web of Science.

This review is organized in four main sections: the first section dis-
cusses bare ceria systems, in the second section ceria-based solid
solutions are reviewed, while the third part covers other metal/
metal oxide doped cerias; finally, before the concluding section,
soot oxidation in an O2/NOx atmosphere is summarized.

11.3 Bare Ceria Formulations


The ability of ceria to promote the oxidation of light and heavy
hydrocarbons is well known.77 Both surface and lattice oxygen are
involved in promoting oxidation reactions, the latter being impor-
tant in assisting oxidation of large hydrocarbons adsorbed on the
surface of the ceria when contact with gas-phase oxygen can be
more difficult. This resembles the oxidation of soot particles in con-
tact with a catalyst.56 For these reasons several studies have been
carried out in the last decade showing the potential of ceria in the
oxidation of carbon particles. However, due to the extreme variety
of conditions employed, quantification of the oxidation potential of
ceria as a function of the structural and morphological parameters

b1469_Ch-11.indd 575 4/8/2013 12:39:05 PM


b1469 Catalysis by Ceria and Related Materials

576 E. Aneggi, C. de Leitenburg and A. Trovarelli

350

300

∆Tm (°C) (Tmsoot-Tmsoot/ceria)


250

200

150

100

50

0
0 5 10 15 20 25 30 35 40 45 50 55
Number of samples

Figure 11.4 Values of ∆Tm (difference between the temperature of oxidation of


soot in the catalysed and uncatalysed reactions) measured over bare ceria (analysis
of literature data).

of the materials employed is very difficult. Regardless of the prepara-


tion method employed (precipitation, thermal decomposition,
inverse micro-emulsion or combustion), in all cases ceria positively
influences soot oxidation temperatures: Fig. 11.4 shows that, exclud-
ing a few cases, the majority of investigations found that ceria can
decrease the oxidation temperature of carbon soot particles by
more than 100 °C, with several samples showing a ∆Tm (difference
between Tm in the presence and absence of a catalyst) between 200
and 300 °C. The great variability of the results mainly reflects the
experimental conditions employed and the different textural or
morphological properties of the oxide due to the preparation and
treatment approaches. Great care should therefore be used to com-
pare results from different studies.
The effect of crystallite size has been investigated by Kockrick
et al.,78 who compared soot combustion over cerium dioxide nano-
particles prepared by an inverse micro-emulsion synthesis and by a
simple precipitation route. The results show that, besides there
being a great difference in crystal size, the catalytic activity for fresh

b1469_Ch-11.indd 576 4/8/2013 12:39:05 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Formulations for Catalysts for Diesel Soot Combustion 577

520

500

Tm (°C) 480

460

440

420

400
0 5 10 15 20 25
Crystal size (nm)

Figure 11.5 Crystal size dependence of catalytic activity (elaboration of the results in
Kockrick et al.78).   samples calcined at 400°C. c samples calcined at 600 °C. Open
symbol for samples prepared by precipitation. Reprinted with permission from Elsevier.

samples is comparable in the two samples and the activity follows the
order of crystal size only within a homogeneous series of samples
(Fig. 11.5). Other effects, like agglomeration of ceria particles, can
mask the effect of crystal size dependence by lowering the catalyti-
cally active surface and leaving crystal size unaltered.
In addition to crystal size, the effect of surface area (SA) and
porosity has also been investigated. It is generally shown that a
reduction in surface area reduces Tm; this is especially true if ceria
samples of the same origin are investigated (e.g. samples of ceria
exposed to different aging temperatures).57,62,64,67,79 Rather scattered
results are found when directly comparing oxidation temperature
with surface area from data obtained in different laboratories
(Fig. 11.6). Although a general trend shows that the higher the
surface area the lower the oxidation temperature, the effect of
experimental set-up and preparation method overlap and mask the
effect of surface area.
Palmisano et al.79 studied the influence of microstructure on soot
activity. Ceria prepared by solution combustion synthesis (SCS) and

b1469_Ch-11.indd 577 4/8/2013 12:39:06 PM


b1469 Catalysis by Ceria and Related Materials

578 E. Aneggi, C. de Leitenburg and A. Trovarelli

600

Tm (°C) 500

400

300

200
0 20 40 60 80 100
2
Surface area (m /g)

Figure 11.6 Surface area dependence of catalytic activity (elaboration of literature


results).

by simple thermal decomposition of cerium nitrate were compared.


The two samples, though having the same surface area (around
60 m2.g−1), were characterized by different pore size distributions;
the SCS ceria exhibited a monomodal size distribution in the range
5–15 nm, while ceria prepared by nitrate decomposition was charac-
terized by a multimodal pore size distribution, with peaks at about
60, 15 and 2 nm. The authors attributed the greater activity of SCS
ceria to a microstructure that increases the soot–catalyst contact
points and maximizes the release of oxygen.
The opposite results were found by Aneggi et al.,80 who studied
the influence of pore size distribution on three samples of ceria,
with the same surface area but prepared by different methods
(Table 11.2). CA was a commercial sample prepared by a proprie-
tary method and calcined at 773 K. CB was prepared according to a
process that involved the precipitation of a homogeneous acidic
solution of cerium nitrate with a base. The CC sample was freshly
prepared by moderate calcination (473 K) of cerium carbonate
obtained from Ce(NO3)3 by precipitation with NH4HCO3. Samples
CB and CC have a smaller pore size (5–6 nm) than CA (14 nm) and

b1469_Ch-11.indd 578 4/8/2013 12:39:06 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Formulations for Catalysts for Diesel Soot Combustion 579

Table 11.2 Catalytic activity results and average pore size distribution.80
Average pore Cumulative pore
Sample Tm(°C) SA(m2.g−1) size (nm) volume (cm3.g−1)
CeO2 (CA) 387 44 14 0.126
CeO2 (CB) 434 42 5 0.049
CeO2 (CC) 440 40 6 0.046

they have a higher oxidation temperature. This was probably due to


the optimization of the soot/catalyst contact, which was closer to the
dimensions of the pores and the soot particles (Printex-U with par-
ticle size of the order of 25 nm).
A large number of investigations have studied the effect of
ceria/soot contact. Issa et al.21,81 performed very detailed studies on
the effect of the contact between soot and catalyst by varying the
type of contact (tight and loose) and the soot/catalyst ratio. They
found that activity increases with better contact between the soot
and the catalyst and that a high Ce/soot ratio increases activity due
to a closer contact between the two solids. They also investigated the
positive effect of increasing the oxygen concentration in the gas
phase with the aim of modelling the kinetics of carbon oxidation82
and pointed out that in ceria-catalysed soot oxidation both gaseous
and ceria lattice oxygen are involved in the reaction.
Saab et al.22,24 investigated the contact points between ceria and
soot by electron paramagnetic resonance (EPR) and found that
under tight contact conditions, a new paramagnetic species is
formed, which acts as a fingerprint of the soot and catalyst interface.
This species is responsible for the increase of activity of soot–catalyst
mixtures.
Using environmental transmission electron microscopy (ETEM)
Simonsen et al.23 showed the evolution of ceria-soot contact points.
They showed how the soot particles in contact with CeO2 are
progressively consumed during oxidation. Transmission electron
microscopy revealed that soot particles are oxidized in close prox-
imity to the interface but the contacts between ceria and soot are
maintained during soot oxidation because they are continuously

b1469_Ch-11.indd 579 4/8/2013 12:39:06 PM


b1469 Catalysis by Ceria and Related Materials

580 E. Aneggi, C. de Leitenburg and A. Trovarelli

re-established during the reaction via some kind of attractive van


der Waals forces between the soot and ceria. In close agreement
with this study, by studying the amount of oxygen transferred from
ceria to soot over several TPO cycles in the absence of oxygen,
Bassou et al.83 found that ceria–soot contact is constant during the
progressive oxidation of soot. They found that the amount of oxy-
gen transferred from the ceria to the soot (and linked to the num-
ber of contact points between the ceria and soot) is constant with
progressive soot oxidation regardless of the type of ceria–soot mix-
ture (i.e. loose or tight) and the ceria/soot weight ratio. In a follow-
ing study, they used a microkinetic approach to understand the key
kinetic parameters governing the catalytic reaction.84 They sug-
gested that an active surface oxygen species of ceria (Oasce) formed
at the interface with the soot. They proposed two different ceria
oxygen reservoirs, which could participate in the reaction at differ-
ent temperature ranges. The first reservoir is constituted by the Oasce
species, the other is represented by surface (Osce) and bulk (Obce)
oxygen not in contact with the soot surface. Osce and Obce may dif-
fuse on the soot/ceria interface, forming additional Oasce at higher
temperatures.
Regarding the mechanism, several authors have proposed a
redox route for soot oxidation,22,24,57,85,86 which utilizes some form of
oxygen from the support in a typical reduction/oxidation mecha-
nism where the catalyst undergoes a partial reduction. An alternative
redox route involves oxygen from the support and gas phase, which
reacts with soot to form adsorbed CO2 in the form of surface carbon
oxygen complexes (like adsorbed carbonates). Decomposition of
carbonate is then stimulated by gas-phase oxygen, which also re-
oxidizes the support, as illustrated below:58

Mred + gas–O → Mox-O (11.1)


Mox-O + Cf → Mred + SOC (11.2)
SOC → CO2 or CO (11.3)

where Mred and Mox–O represent the reduced and oxidized state of
the ceria, respectively. Gas–O indicates the oxidant gas (O2, NO,

b1469_Ch-11.indd 580 4/8/2013 12:39:06 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Formulations for Catalysts for Diesel Soot Combustion 581

NO2, etc.), Cf the carbon active site on the carbon surface and SOC
a surface carbon–oxygen complex. In the first equation ceria is oxi-
dized by the gas phase, in the second step the oxygen of the catalyst
is transferred to the active site on the carbon surface reducing the
catalyst and then the surface carbon–oxygen complex is decom-
posed to CO and CO2. In addition to a classical redox route a
mechanism that relies on the presence of active oxygen species has
also been proposed.87 In this case the role of the ceria could be to
cause spillover of active oxygen onto the soot surface, with active
oxygen species likely existing in the form of peroxide or superox-
ide.64,87,88 Active oxygen like superoxide (O2−) could form via oxygen
adsorption on the reduced surface of ceria in the vicinity of soot
according to:

Ce3+− + O2 → Ce4+–(O2)−

which then reacts with carbon to form CO2:

2[Ce4+−(O2)−] + C → 2[Ce4+−(O)] + CO2

In this case reduced ceria would be the starting point for the forma-
tion of active superoxide ions. Ceria can be reduced by soot at the
contact points located at the soot/ceria interface and this will initi-
ate the ‘active oxygen’ path, which would contribute to the direct
oxidation of soot particles by a spillover mechanism. The redox of
ceria is therefore important in the generation of reduced ceria at
the interface points and lowering the reduction temperature will
therefore increase the overall activity of soot oxidation.
This indicates that in order to improve the soot oxidation abili-
ties of ceria efficiently it is important to modify its redox capacity
and to enhance the formation of active oxygen species like perox-
ide or superoxide. This can be conveniently done either through
modification of the structural features of ceria by lattice doping
(i.e. with rare-earth elements or zirconium) or by doping with met-
als able to induce the formation of active oxygen species like silver
(see later).

b1469_Ch-11.indd 581 4/8/2013 12:39:06 PM


b1469 Catalysis by Ceria and Related Materials

582 E. Aneggi, C. de Leitenburg and A. Trovarelli

11.4 Ceria-Based Solid Solutions


The introduction of isovalent and aliovalent elements into a ceria
lattice can induce important modifications of its textural, structural
and redox properties and this can deeply affect its catalytic proper-
ties.89–91 Ceria–zirconia still represents the best example showing
how the incorporation of an isovalent element like Zr into the
ceria structure can enhance the thermal, redox and catalytic
properties.92–94 The state-of-the-art material for oxygen storage in
TWC is in fact appropriate amounts of zirconia-modified ceria
mixed oxides including small amounts of other rare-earth elements,
which are added to further enhance the redox and textural features
of ceria.34,38,95
The introduction of Zr4+ (and in general the substitution of Ce4+
for other transition metals and rare-earth elements) induces struc-
tural modifications into the fluorite lattice of CeO2,33,41,42,95 which is
attributable to the smaller dimension of the Zr4+ ion (0.84 vs.
0.97 Å96). This results in a decrease of cell volume and the formation
of structural defects in the lattice, increasing the rate of oxygen dif-
fusion through the material. Depending on preparation conditions
and the concentration of the single-oxide constituents, several
phases can form and coexist in the CeO2–ZrO2 system.97,98 In gen-
eral, a single-phase monoclinic cell is observed for a CeO2 content
less than 20 mol%; in CeO2-rich compositions (CeO2 > 70 mol%)
solid solutions of cubic symmetry are formed. At intermediate levels,
regions of tetragonal and cubic symmetry coexist in the phase dia-
gram, their formation depending on the preparation method used.
In the intermediate region it is possible to find several tetragonal
phases: the t phase (stable), the t’ phase (metastable with c/a > 1,
where c/a is the axial ratio and a and c are the two side of the tetrag-
onal structure) and the t’’ phase (metastable, with c/a = 1).98 The t’’
phase is a pseudo-cubic phase where oxygen anions are shifted from
the ideal position of the fluorite structure. The phase composition
can also severely influence the redox behaviour of cerium99 because
oxygen mobility is generally favoured in the cubic structure, whereas
the tetragonal structure is less active in promoting oxygen mobility.

b1469_Ch-11.indd 582 4/8/2013 12:39:06 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Formulations for Catalysts for Diesel Soot Combustion 583

A better synergism of all these factors can be achieved with ternary


oxides,100–103 based mainly on CeO2–ZrO2–MOx (M = Pr, Y, La, …),
which can give near-optimal reactivity for oxygen storage and release
when coupled with higher thermal resistance.
The modification of ceria with other rare-earth and transition-
metal elements has been one of the approaches used to enhance the
soot oxidation ability of ceria. Modified ceria catalysts lead to
improved soot oxidation activity and have been investigated in great
detail.58,61,64,66,104–108 However, the scarceness of systematic studies in
the literature does not allow the creation of a rational picture of the
role of the dopant in modifying the textural, morphological, redox
and catalytic activity of CeO2 in soot oxidation. Certainly the choice
of dopant and the synthesis of materials play a major role (the for-
mation of a homogeneous solid solution or mixed phase oxides is
generally dependent on this step) but critical variables are also sur-
face area and the pore distribution system. As reported in the previ-
ous section, the generally accepted mechanism for the catalytic
oxidation of soot is a redox mechanism in which the metal site is
oxidized and reduced consecutively. Several authors agree that the
reducibility and redox activity of the samples can be correlated to
soot oxidation activity.58,106,107,109,110 It has been well established that
doping ceria with Zr results in more active soot oxidation catalysts;
however, composition, surface area, pore structure, preparation and
thermal treatments can deeply affect the extent of the effect of
zirconium.66,108,109,111,112 We found that activity is mainly related to two
factors: the oxygen storage and redox capacity, and surface area.57,62
Regarding surface area, a strong dependence is observable for
low-surface-area samples, while activity is less influenced when the
final surface areas are higher than 40 m2.g−1. More important is oxy-
gen storage capacity (OSC), which provides an alternative route for
the oxidation of large soot particles in contact with the ceria.56,57,62
The results indicate a strong relation between catalytic activity and
OSC, pointing out that the participation of lattice oxygen in soot
oxidation may contribute to overall activity, especially when the sur-
face area is low or in the presence of large soot particles, when
access to gas-phase oxygen is limited by geometrical factors. Similar

b1469_Ch-11.indd 583 4/8/2013 12:39:06 PM


b1469 Catalysis by Ceria and Related Materials

584 E. Aneggi, C. de Leitenburg and A. Trovarelli

600 450

400

350

O2 extracted (µmol/g)
500 300
Tm (°C)

250

200

400 150

100

50

300 0
20 30 40 50 60 70 80 90 100
Amount of ceria (mol %)

Figure 11.7 Tm (filled symbols) and OSC (open symbols) against composition.
(Square symbols (tight contact) from Aneggi et al.;57,62 circles (loose contact) from
Zhu et al.67). Reprinted with permission from Elsevier.

conclusions were reached by Zhu et al. for a series of CexZr1−xO2 cata-


lysts investigated under loose contact conditions.67 The oxidation
temperature for soot shows a strong dependence on the composi-
tion of the solid solution and, under both loose and tight condi-
tions, a minimum of the oxidation temperature was found in the
middle composition range (Fig. 11.7). Furthermore, for both series
of materials, the OSC profile is perfectly symmetrical with catalytic
activity indicating the contribution of the lattice oxygen in the reac-
tion. The role of lattice oxygen and reducibility was also evidenced
by Atribak et al.,109 who studied the effect of thermal treatment in
soot oxidation over samples of ceria and ceria–zirconia calcined at a
high temperature (1000 °C).
In several studies the key roles of surface area and oxygen stor-
age capacity have been taken into account but their relative
importance is still a matter of discussion. In a recent study we
investigated in more detail the activity of a series of ceria–zirconia-
based catalysts.111 Ceria–zirconia solid solutions over the whole

b1469_Ch-11.indd 584 4/8/2013 12:39:06 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Formulations for Catalysts for Diesel Soot Combustion 585

composition range were prepared by co-precipitation and the


oxidation of soot in the temperature range 300 < T < 600 K was
systematically investigated paying special attention to the depend-
ence of activity on surface area and redox behaviour. The surface
area of all samples varied in the range 10-100 m2.g−1. For all calci-
nation temperatures, the Ce0,75Zr0,25O2 composition was the most
active due to an appropriate combination of surface area and
surface cerium active sites. In this study it was clearly evidenced
that in the presence of gas-phase oxygen, soot oxidation activity
depends on the number of available surface oxygen species
(which depends on surface Ce loading, which, in turn, should be
dependent on the nominal composition excluding important sur-
face segregation phenomena) while scarce or no dependence was
observed between oxidation activity and oxygen storage capacity,
thus excluding the significant participation of lattice oxygen in
the reaction. An almost opposite picture was observed when oxi-
dation was performed in the absence of gas-phase oxygen, when a
strong correlation between oxidation activity and OSC was clearly
evident (Fig. 11.8).
Therefore, in order to enhance activity, it is very important to
maximize the number of active surface oxygen species, but oxygen
storage is important because it provides an alternative route for the
oxidation of large soot particles in contact with ceria or it can pro-
vide oxygen for oxidation in a rich environment.
The substitution of Ce in CeO2 by Hf results in the formation
of ceria–hafnia solid solutions, whose properties in soot oxidation
were investigated by Reddy et al.113 The order of soot oxidation
activity (Ce–Hf > Ce–Zr) was correlated to the number of oxygen
vacancy defects as evidenced by Raman studies. Moreover, diffuse
reflectance ultraviolet-vis spectroscopy patterns showed that the
ceria–hafnia sample had the best resolved band at about 255 nm,
which has been characterized as a Ce3+ ← O2− charge transfer tran-
sition and correlated to a higher concentration of oxygen vacancy
defects. In a recent paper the same group also compared the cata-
lytic activity of pure ceria, ceria–zirconia and ceria–lanthana solid

b1469_Ch-11.indd 585 4/8/2013 12:39:06 PM


b1469 Catalysis by Ceria and Related Materials

586 E. Aneggi, C. de Leitenburg and A. Trovarelli

800 3,5

780
3
760
740
2,5

Weight loss (%)


720
T50 (K)

700 2

680
1,5
660
640
1
620
600 0,5
0 40 80 120 160 200 240 280
0

500

1000

1500

2000

2500

3000

3500
Total surface available oxygens
(µmol O/g) OSC (µg O2/g)

(a) (b)

Figure 11.8 (a) T50 against total surface oxygen availability (from TG experiments
on soot/catalyst mixtures under air). (b) Oxidation in the absence of gas-phase
oxygen against OSC (from TG experiments on soot/catalyst mixtures under inert
atmosphere).111 Reprinted with permission from Elsevier.

solutions pointing out that doped ceria showed higher OSC, better
thermal stability and enhanced activity.114 They used several experi-
mental techniques to investigate the reaction mechanisms, con-
cluding that the introduction of La3+ into the ceria structure
induces a lattice deformation creating oxygen vacancies, which are
responsible for the higher activity in comparison to ceria–zirconia,
although the different Zr and La loadings might have affected
their conclusions. The proposed mechanism involved the forma-
tion of active oxygen by the filling of vacancies with gaseous oxygen
followed by the transfer or spillover to soot particles.
The improvement of soot oxidation activity resulting from the
introduction of La3+ into CeO2 was also reported by Bueno-López
and co-workers;58 they studied the influence of doping ceria with
different amounts of La3+ and correlated the increase of activity to
an enhancement of the surface area of CeO2 due to the La3+ and
to an improvement in the redox properties. The incorporation of

b1469_Ch-11.indd 586 4/8/2013 12:39:06 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Formulations for Catalysts for Diesel Soot Combustion 587

La3+ was shown to improve the generation of active oxygen species,


which react with carbon sites. The improvement of activity related
to La3+ doping was seen only under tight contact whereas under
loose contact the activity was dependent only on surface area. No
improvement of activity was seen when using mixtures of La2O3
and CeO2, thus pointing out the importance of the formation of a
solid solution. In another study, La2O3 was used as a surface pro-
moter of CeO2 and a mechanism involving a carbonate route and
associated with the basicity of La2O3 was proposed to explain the
results.61
The incorporation of yttrium in ceria resulted in catalysts that
were more active under tight contact, while no difference was found
under loose contact.104 The benefits of the incorporation of yttrium
were visible only at low loading and were related to the surface seg-
regation of Y3+, which promoted the creation of oxygen vacancies
and hindered crystallite growth. The incorporation of Y3+ in ceria–
zirconia did not result in more active catalysts due to the prevailing
effect of zirconium.
Fang et al.106 studied solid solution catalysts of ceria doped with
Zr, Pr and Gd and found that surface area is not a crucial parameter;
an analysis of H2 consumption in temperature-programmed reduc-
tion suggests that the overall activity follows the reducibility of ceria
in solid solutions (Fig. 11.9) although reducibility alone cannot
totally account for the observed behaviour.
Masui et al.,115 studying a ceria–zirconia solid solution doped
with a small amount of Bi2O3, demonstrated that the active oxygen
released from the catalyst correlates with soot oxidation. The intro-
duction of Bi2O3 into the fluorite structure increased the oxygen
storage capacity of the system and decreased the oxidation tem-
perature to about 60 °C compared with undoped ceria–zirconia.
The OSC of the Ce0.69Zr0.11Bi0.20O1.9 catalyst was more than twice
that of a conventional Ce0.86Zr0.14O2 material; the reason for such a
high OSC was that Bi2O3 is easily reduced to metallic Bi and that
both Ce4+ and Bi3+ are reduced simultaneously. Recently, the
same group found very good activity for a ternary solid solution of
CeO2–Pr6O11–Bi2O3 due to enhanced redox activity induced by the

b1469_Ch-11.indd 587 4/8/2013 12:39:06 PM


b1469 Catalysis by Ceria and Related Materials

588 E. Aneggi, C. de Leitenburg and A. Trovarelli

65 Ce0.7Zr0.3O2 4,5

Ce0.7Pr0.3O2-δ 4
55 3,5

H2 consumption ratio
Surface area (m2/g)
Ce0.7Pr0.3O2-δ 3
45
Ce0.7Gd0.3O2-δ
2,5
2
35 Ce0.7Zr0.3O2
Ce0.7Gd0.3O2-δ
1,5

25
CeO2 1
0,5
CeO2
15 0
380 400 420 440 460 480 500
Tm (°C)

Figure 11.9 Surface area ( ) and H2 consumption ratio (•) against Tm.106
Reprinted with permission from Elsevier.

transition from Pr4+ to Pr3+.116 The formation of a CeO2-Pr6O11-Bi2O3


solid solution promotes the release of active oxygen at low tempera-
tures and improves soot oxidation, the most active system having a
T50 of 349 °C.
The introduction of copper into ceria and ceria–zirconia (only
partially in the form of solid solutions but mainly dispersed in the
surface and sub-surface regions) also showed benefits in terms of a
lower operating temperature.117,118 The enhancement of activity was
attributed to dispersed surface and sub-surface Cu species in close
interaction with ceria and ceria–zirconia, which promoted the for-
mation of highly reactive and active oxygen species, previously
described by Bueno-López et al.58 Similarly, active oxygen in
Fe-doped ceria was found to be responsible for an enhancement in
the activity of ceria when Fe enters into the solid solution.62,110
A redox mechanism that involves cycling of Ce4+/Ce3+ and Fe3+/Fe2+
has been proposed to account for the observed behaviour. The close
similarities between the mechanisms proposed in these investiga-
tions can be summarized schematically in a redox cycle involving an

b1469_Ch-11.indd 588 4/8/2013 12:39:06 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Formulations for Catalysts for Diesel Soot Combustion 589

M–O–Ce species, where M is an aliovalent (M+/2+/3+) transition metal


inserted into CeO2 lattice:

O
n+ 4+
(O-) M - O - Ce - O -
(n-2) gas/bulk O2
O O

C O
n+ 3+
COx (O-) M - O - Ce - O -
(n-2)
O
O
(n-1)+ 4+
(O-) M - O - Ce - O -
(n-2)
O

The oxygen anion bound to the metal site reacts with carbon
moiety to give COx with the generation of an oxygen vacancy and a
reduced metal site. This is then refilled by oxygen from an adjacent
Ce4+ site, which is concomitantly reduced to Ce3+. At this point the
active oxygen species is reformed and the Ce3+ site can be oxidized
by gas-phase oxygen or (under oxygen-deficient atmospheres) by
oxygen diffusion through the bulk.
In addition to the textural and redox properties, pore structure
is also important for soot oxidation activity. Krishna et al. studied
ceria doped with La, Pr, Sm and Y.66,105 They found that the forma-
tion of macropores, which increase the accessible surface area, mul-
tiplies the contact between the catalyst and soot thus increasing
activity (Table 11.3). Indeed, by doping ceria with Pr3+/4+ and La3+,
the pore size distribution assumes a bimodal profile and is shifted
towards mesopores and macropores, increasing the external surface
area and promoting activity. In agreement with this investigation,
Aneggi et al. compared pore size distributions for a series of ceria–
zirconia solid solutions.111 When considering samples with similar
surface area, independently of the composition, a decrease in pore
size induces a decrease in activity. This is true until the surface area
is maintained above a certain minimum value; when the surface
area drops to lower values, no correlation between activity and pore
size distribution could be found.

b1469_Ch-11.indd 589 4/8/2013 12:39:07 PM


b1469 Catalysis by Ceria and Related Materials

590 E. Aneggi, C. de Leitenburg and A. Trovarelli

Table 11.3 Catalytic activity and structural properties of samples.


Sample Tm (°C) Pore size (nm) SA (m2.g−1) Reference
CeO2_aged 700 450 10 57 105
CeLaOx_700 425 15; 50 48 105
CePrOx_700 422 15; 50 44 105
CeO2_500 378 20 53 111
Ce0.75Zr0.25O2_600 380 20 58 111
Ce0.44Zr0.56O2_700 406 15 53 111
Ce0.25Zr0.75O2_700 422 9 59 111

Recently catalytic activity in soot oxidation of three-dimensionally


ordered macroporous (3DOM) materials with a uniform pore size
was investigated.119 Ordered ceria–zirconia and doped ceria–zirconia
solid solutions were tested and the activity of the 3DOM materials
was higher compared to disordered macroporous samples, indicat-
ing the importance of an open periodic three-dimensional frame-
work in improving soot catalyst contact efficacy and thus catalytic
activity. Interestingly, the 3DOM samples were characterized by
lower surface area and larger crystallites than disordered macropo-
rous materials and no relation was found between oxidation tem-
perature and crystallite size or surface area. This indicates the
critical aspects of channel structure and interconnections between
pores, in addition to the pore volume itself.
Overall, the activity of ceria for soot oxidation can be improved
by doping with other transition or rare-earth elements. It seems that
the optimum benefit is reached when doping causes the formation
of solid solutions, at least in the surface or sub-surface regions. In
terms of oxidation performance, the characteristic temperature for
the oxidation of the most active samples is in the range 370–390 °C
(see Table 11.4 for a summary).

11.5 Formulations Based on Metal-Doped Ceria


The synergetic interaction between ceria and other metals, when solid
solutions are not formed, can also result in very active and resistant
catalysts. Several catalyst formulations containing supported transition

b1469_Ch-11.indd 590 4/8/2013 12:39:07 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Formulations for Catalysts for Diesel Soot Combustion 591

Table 11.4 Soot oxidation temperatures (Tm) over ceria–zirconia ceria-based solid
solutions.
Aging Soot/catalyst
temperature ratio (type
Sample (°C) of contact) O2 (%) Tm (°C) Reference
Ce0.75Zr0.25O2 500 1:10 (tight) 10 380 112
CeO2+3%La 500 1:20 (tight) 20 383 61
Ce0.75Zr0.25O2 500 1:20 (tight) 6 387 62
CeO2 500 1:20 (tight) 6 389 62
Ce0.70Zr0.30O2 600 1:9 (tight) 20 397 106
Ce0.75Zr0.25O2 750 1:20 (tight) 6 400 62
Ce0.44Zr0.56O2 750 1:20 (tight) 6 401 62
CeO2+3%La 750 1:20 (tight) 20 408 61
Ce0.8Hf0.2O2 500 1:4 (tight) 20 409 113
CeO2–5La 1000 1:4 (tight) 20 410 58
Ce0.70Pr0.30O2−δ 600 1:9 (tight) 20 414 106
Ce0.90Pr0.10O2−δ 700 1:4 (tight) 20 422 105
Ce0.90La0.10O2−δ 700 1:4 (tight) 20 425 105
Ce0.69Zr0.31O2 1000 1:4 (tight) 20 455 109
Ce0.84Zr0.15Y0.01O2 1000 1:4 (tight) 20 456 104
Ce0.50Zr0.50O2 700 1:10 (loose) 10 525 67
Ce0.75Zr0.25O2 700 1:4 (loose) 5 585 108

and noble metals or alkali metal elements have been used with suc-
cess in soot oxidation.76,117,120–128 The promotion of ceria with alkali
metals60,63,68,129,130 (and especially potassium) results in a dramatic
decrease of the temperature of soot combustion, with oxidation
temperatures in the range 350–400 °C. This activity enhancement is
dependent on metal loading; in the case of potassium, compositions
containing 7–10 wt% of alkali metal are associated with optimal
activity60,129 (Fig. 11.10). Several mechanistic studies were carried out
in order to understand the role of potassium and it is generally
accepted that the high mobility of alkali metal species plays a key
role in the improvement of catalytic activity of alkali metal com-
pounds during soot oxidation. However, this is not the only reason
for activity enhancement.

b1469_Ch-11.indd 591 4/8/2013 12:39:07 PM


b1469 Catalysis by Ceria and Related Materials

592 E. Aneggi, C. de Leitenburg and A. Trovarelli

500

450
Tm (°C)

400

350

300
0 2 4 6 8 10 12 14
amount of K (%)

Figure 11.10 TPO results for CeO2 doped with K (y from Aneggi et al.60, „ from
Gross et al.129). Reprinted with permission from Elsevier.

The activity of K-supported catalysts was investigated by Wu


et al.63 on potassium nitrate impregnated on CeO2 and Ce0.5Zr0.5O2.
The influence of KNO3 on activity was attributed to several causes:
(i) a liquid phase generated by the melting of potassium nitrate,
which increases the soot–catalyst contact; (ii) alkali metal–carbon
intermediates, which favour gasification and (iii) the oxidizing
potential of nitrate ions, which could act as oxidizing agents being
reduced to nitrite ions. The effect of potassium on the redox of
cerium was not investigated in this study. Different mechanisms
that involve the interaction of alkali metals with the redox couple
of ceria through the formation of carbonate species were recently
proposed.60,68,129 A mechanism based on a synergetic effect between
ceria and potassium was suggested by Peralta et al.68 with ceria pro-
viding active oxygen thanks to its redox capacity, and K acting by
forming carbonate-type intermediates with the partially oxidized
soot, which then decomposes and releases CO2. The high mobility
of K compounds, due to their relatively low boiling points, may also
improve the effective contact between the active phases and soot.

b1469_Ch-11.indd 592 4/8/2013 12:39:07 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Formulations for Catalysts for Diesel Soot Combustion 593

A detailed elaboration of the previous mechanism was later sug-


gested by the same group,129 who proposed a mechanism for soot
oxidation that includes the reduction and oxidation of ceria and
the formation of potassium carbonate, which decomposes to CO2.
A key step in the proposed mechanism was the formation of perox-
ide or superoxide species on the surface of the ceria, which react
with soot promoting the transformation of KNO3 to K2CO3 and
nitrites.
We also investigated the role of alkali metals on ceria and other
supports, using different alkali metal precursors.60 We found that
the enhancement of activity was not correlated to the nature of the
support but was dependent on the nature and loading of the ele-
ment (Cs being more active than Rb ≈ K > Na) and on the type of
precursor salt used. The activity enhancement can be interpreted in
terms of an oxygen exchange mechanism in which carbon–oxygen
complexes are the active sites, in a situation that resembles those
found in potassium-promoted gasification of carbonaceous mate-
rial.131,132 The effect of alkali metals is to favour the chemisorption of
molecular oxygen with the formation of carbon surface complexes
that eventually react with the soot. Consequently the catalyst is
an oxygen carrier that transfers oxygen from the gas phase to the
carbon surface.
This mechanism, which does not rely on the redox of ceria, is in
agreement with the fact that the action of alkali metals seems to be
independent of the type of support. Milt et al.133 found for example
that potassium promotes the activity of La2O3 thus decreasing the
temperature of soot oxidation to the range 350–400 °C and attrib-
uted the increase of activity to the dynamic of formation and the
decomposition of carbonate-type intermediates.
Potassium acting with a redox mechanism correlated to the
redox cycle of cerium oxide is also a possibility.60 In this case the
participation of potassium can be connected to the oxidation/
reduction cycle depicted in the scheme below:
C CeO2-j+z KxOy
O2
CO2 CeO2-j KxOy+1

b1469_Ch-11.indd 593 4/8/2013 12:39:07 PM


b1469 Catalysis by Ceria and Related Materials

594 E. Aneggi, C. de Leitenburg and A. Trovarelli

KxOy is oxidized by O2 to give KxOy+1,134,135 which transfers oxygen to


reduced ceria, which can then reduce soot particles.
One of the drawbacks of the systems containing alkali metals is
their resistance under hydrothermal conditions and the presence of
sulfur.60,68 In the first case the activity of soot oxidation drops dra-
matically for K/CeO2 catalysts under water at higher temperatures
due to the loss of potassium and a change in the K/Ce optimal ratio,
while the formation of sulfates is responsible for activity loss
under SO2.
Of the transition metals, Mn, Cu and Co have often been
employed as dopants and promoters of soot oxidation catalysts
based on ceria. Manganese oxides, MnO2, Mn3O4 and Mn2O3,
absorb oxygen under oxidizing conditions and are reported to be
active and stable oxidation catalysts for the combustion of organic
compounds.136,137 The doping of ceria with manganese also resulted
in active and stable catalytic materials in the oxidation of organic
compounds138 and the recent use of Mn as a dopant in ceria-based
soot oxidation catalysts confirms this tendency. Irrespective of the
metal addition method employed, whether wet impregnation, the
sol-gel method or the complexation-combustion process, an
enhancement of activity is observed.117,122,123,139
Saab et al. found that doping ceria with manganese leads to a
very active catalyst compared to Mn/Al2O3, and the activity was cor-
related to the presence of Mn species in a high oxidation state.123
EPR studies, indeed, showed that some manganese species with
higher oxidation states, in the presence of soot, are reduced to Mn2+
ions, which can be considered the first step in the soot combustion
mechanism. The same effect was not observed over Mn/Al2O3.
Similarly, Mn deposited over ceria–zirconia decreased the combus-
tion temperature of a sample of real diesel soot to the range
300–375 °C, although in this case, due to the nature of the soot,
oxidation required more than one step.122
Liang et al.117 studied the incorporation of Mn into CeO2 to form
a solid solution. The incorporation of Mn was reported to increase
the generation of anionic vacancies, which enhanced the adsorption

b1469_Ch-11.indd 594 4/8/2013 12:39:07 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Formulations for Catalysts for Diesel Soot Combustion 595

of active oxygen species. In this case, however, the variation of the


cell parameter due to the incorporation of Mn in the fluorite struc-
ture was too low to be ascribed to the formation of a solid solution
when compared with the value obtained according to Vegard’s law140
(5.32 Å for Mn4+ and 5.23 Å for Mn3+ against the reported value of
5.406 Å). It is likely that manganese oxide well dispersed on the ceria
surface or in an amorphous state coexists with manganese located in
the fluorite structure. The structural characterization revealed the
presence of at least two Mn species in the sample: a fraction of Mn
was incorporated into the ceria lattice and the remainder was homo-
geneously distributed on the surface of the catalyst.
Several papers consider the promotional effects of copper on
ceria-based catalysts,117,118,141–145 which results in very active soot com-
bustion with a temperature for soot oxidation around 350 °C. The
commonly proposed mechanism for the reaction is correlated with
the presence of CuO particles and their capability of being reduced
to Cu+ and then re-oxidized to Cu2+. The reduction of supported
copper species by carbon during soot combustion was confirmed by
EPR analysis by Lamonier et al.142,146 Liang and co-workers consid-
ered the copper oxide–ceria interface suggesting that the pairs
Ce4+-O2−–Cu2+ could be rapidly reduced by soot to produce CO2.117,118
After the reduction of copper species and the combustion of soot,
the ceria lattice supplies the lost oxygen anion to the copper–ceria
interface and the vacancies formed during the reaction are refilled
by gaseous oxygen. The beneficial effect of CuO was also demon-
strated for a ceria–zirconia solid solution doped with cobalt and
nickel.141 Adding copper to these catalysts increased the activity by
more than 60 °C, lowering the oxidation temperature of soot to
below 400 °C.
Catalyst formulations containing ceria doped with cobalt have
also been tested.47,76,147 Cobalt is generally present on the surface of
ceria as Co3O4,76,147 and the activity is associated with its reduction to
CoO assisted by oxygen spillover on the ceria support. The redox
activity of ceria can also enhance the reducibility of Co3O4,76,148
which in turn influences its oxidation activity.

b1469_Ch-11.indd 595 4/8/2013 12:39:07 PM


b1469 Catalysis by Ceria and Related Materials

596 E. Aneggi, C. de Leitenburg and A. Trovarelli

A combination of molybdenum with ceria has also shown poten-


tial for soot oxidation.149 At low loading the Mo6+ becomes incorpo-
rated into the surface vacant sites of CeO2 improving catalytic
activity; when the amount of molybdenum increases, the number of
active vacant sites on the ceria surface decreases with a consequent
increase in the temperature of combustion and decrease of selectiv-
ity to CO2.
Enhanced oxidation activity is also found when doping ceria
with noble metals. A moderate increment of activity was observed
with Pt supported by ceria and ceria–lanthana compared to bare
supports,126 while in the presence of Ru an improvement of more
than 100 °C was reported under loose contact conditions.120,121 The
two noble metals act with different reaction mechanisms. Platinum
acts by speeding up the oxygen uptake step, while the high activity
of ruthenium catalysts is due to the weakness of the Ru–O bond,
which increases the mobility of oxygen so that the reduction of Ru
oxide species at low temperatures is easy. Ruthenium takes part in
soot oxidation with its own redox cycles, being reduced to metal by
soot and oxidized by gas-phase oxygen; the coupling of the redox
cycles of RuO2 and CeO2, is likely to favour activity at lower
temperatures.
Silver-based materials are another interesting class of catalysts for
diesel soot oxidation and are among the most promising formula-
tions investigated.150–155 Silver is known to be an efficient oxidation
catalyst and the use of Ag deposited on ceria was also found to
increase the rate of carbon gasification compared to other noble
metals.156 A decrease of the onset oxidation temperature of soot by
more than 50°C was observed on Ag10%/CeO2 compared to
unloaded ceria,87 which has been attributed to the formation of
active oxygen species promoted by silver; it is in fact known that a
metallic silver surface can promote the formation of several suboxide
and superoxide O2− species, which are active in oxidation.87,157–162
Silver nanoparticles are generally considered the active supported
species; Ag2O is reported to be a strong stoichiometric oxidant
for soot but not a catalyst, due to its deactivation after a single
catalytic run.151

b1469_Ch-11.indd 596 4/8/2013 12:39:07 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Formulations for Catalysts for Diesel Soot Combustion 597

Although the activity of Ag/CeO2 catalysts toward soot oxida-


tion is high compared to other silver supported catalysts (with
oxidation temperatures in the range 330 < T < 370 °C), the level of
promotion by silver on ceria is lower than that observed on ZrO2
or Al2O3.150 This behaviour is likely due to a different interaction
between silver and the support, which affects the transformation
Ag2O → Ag during preparation or calcination. In particular zirco-
nia and alumina seem to benefit from the larger quantities of
metallic Ag, while for ceria the larger amount of Ag2O coupled
with the higher initial activity of the bare support reduces the
overall effect of silver. In addition, the dispersion of Ag is also
affected by the support, being higher over alumina and zirconia
and lower with ceria where Ag forms a smooth metallic layer on
top of Ag2O. It is likely that the redox of ceria maintains silver in
an oxidation state at the ceria surface, which is not ideal for
assisting the soot oxidation reaction, whereas when silver itself is
reduced to the zero valent state a more potent soot combustion
catalyst is formed.
Recently, Shimizu and co-workers151,152 studied silver-doped ceria
and found that metallic silver nanoparticles are predominantly
formed on ceria. These enhance the reducibility of ceria with a ben-
eficial effect on catalytic activity. The effect of the preparation and
the deposition method of Ag was shown also to positively influence
catalytic activity. The development of novel preparation methods
was also the focus of the study by Kayama et al.,155 who investigated a
new synthesis route to develop a rice-ball nanostructure (an aggre-
gate of ceria particles with silver at the centre), which is more active
than conventional systems. This approach is directed towards the
maximization of the silver/ceria interface to activate oxygen species,
increasing the contact between the ceria and soot and covering sil-
ver with ceria particles to prevent sintering. Thermal resistance can
also be increased by doping the ceria in the rice-ball structure with
other rare-earth ions.
A summary of the main results showing catalytic activity along
with experimental conditions for metal/metal oxide doped ceria
are summarized in Table 11.5.

b1469_Ch-11.indd 597 4/8/2013 12:39:07 PM


b1469 Catalysis by Ceria and Related Materials

598 E. Aneggi, C. de Leitenburg and A. Trovarelli

Table 11.5 Soot oxidation temperatures (Tm) over metal/metal oxide doped
ceria.
Aging Soot/catalyst
temperature ratio (type of O2 Tm
Sample (°C) contact) soot type (%) (°C) Reference
Co/CeO2 400 1:20 (tight) real 6 300 76
20%Ag/CeO2 500 1:20 (tight) MA7 20 308 152
Cu–Ce 500 1:10 (tight) 9.5 324 118
Printex U
5%Ag/CeO2 500 1:20 (tight) 20 334 150
Printex U
1.5%Ru/CeO2 600 1:4 (tight) N330 20 345 120,121
0.05Mn/Ce 500 1:20 (tight) N330 20 346 123
12%Co,4.5%K/CeO2 400 1:20 (tight) real 6 350 47
7%K/CeO2 450 1:20 (tight) real 6 350 129
1Mo1000Ce 500 1:4 (tight) N330 20 354 149
Cu–Ce (1:9) 500 1:10 (tight) 10 356 117
Printex U
10%K/Ce0.50Zr0.50O2 800 1:10 (tight) 10 359 63
Printex U
KOH(4.5)/CeO2 400 1:5 (tight) real 20 360 130
10%Cs/CeO2 500 1:20 (tight) 6 360 60
Printex U
CuO (5%)–CoO(5%)/ 500 1:4 (tight) 20 363 141
CeO2–ZrO2 Printex U
Cu–Ce–Zr 500 1:10 (tight) 9.5 365 118
Printex U
22%Ba,7%K/CeO2 400 1:20 (tight) real 6 365 68
Mn-Ce (1:9) 500 1:10 (tight) 10 368 117
Printex U
7%K/Ce 400 1:20 (tight) real 6 369 68
10%K/Ce 500 1:20 (tight) 6 370 60
Printex U
10%Cs/Ce 750 1:20 (tight) 6 370 60
Printex U
(Continued )

b1469_Ch-11.indd 598 4/8/2013 12:39:07 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Formulations for Catalysts for Diesel Soot Combustion 599

Table 11.5 (Continued )


Aging Soot/catalyst
temperature ratio (type of O2 Tm
Sample (°C) contact) soot type (%) (°C) Reference
5%Ag/CeO2 750 1:20 (tight) 20 371 150
Printex U
CeO2+4.5%K 400 1:20 (tight) real 6 375 47
CeO2+10%K 600 1:10 (tight) 10 375 63
Printex U
10%K/Ce 750 1:20 (tight) 6 377 60
Printex U
1%Cr/99%CeO2 650 1:9 (tight) 10 377 128
Printex U
10%Rb/Ce 750 1:20 (tight) 6 378 60
Printex U
Cu–Ce 800 1:10 (tight) 9.5 378 118
Printex U
10%Rb/Ce 500 1:20 (tight) 6 380 60
Printex U
Cu–Ce–Zr 800 1:10 (tight) 9.5 384 118
Printex U
CuO (5%)–NiO (5%)/ 500 1:4 (tight) 20 388 141
CeO2–ZrO2 Printex U
5%Ag/Ce 700 1:20 (tight) 20 415 153
CB Sigma
3%Cu/Ce10Zr80Y10O2−δ 250 1:4 (tight) N330 20 421 142,146
10%K/Ce0.50Zr0.50O2 600 1:10 (loose) 10 396 63
Printex U
10%K/CeO2 600 1:10 (loose) 10 398 63
Printex U
10%K/Ce0.50Zr0.50O2 800 1:10 (loose) 10 455 63
Printex U
5%Ag/Ce 500 1:20 (loose) 20 460 150
Printex U
CeO2+5%Ru 600 1:4 (loose) N330 20 490 120,121

(Continued )

b1469_Ch-11.indd 599 4/8/2013 12:39:07 PM


b1469 Catalysis by Ceria and Related Materials

600 E. Aneggi, C. de Leitenburg and A. Trovarelli

Table 11.5 (Continued )


Aging Soot/catalyst
temperature ratio (type of O2 Tm
Sample (°C) contact) soot type (%) (°C) Reference
5%Ag/Ce 750 1:20 (loose) 20 494 150
Printex U
Cu–Ce 500 1:10 (loose) 9.5 496 118
Printex U
Mn–Ce (1:9) 500 1:10 (loose) 10 503 117
Printex U
Cu–Ce–Zr 500 1:10 (loose) 9.5 505 118
Printex U
Cu–Ce 800 1:10 (loose) 9.5 519 118
Printex U
Cu–Ce (1:9) 500 1:10 (loose) 10 522 117
Printex U
Cu–Ce–Zr 800 1:10 (loose) 9.5 526 118
Printex U
5%Ag/Ce 700 1:20 (loose) 20 528 153
CB Sigma
10%K/CeO2 800 1:10 (loose) 10 545 63
Printex U

11.6 Reactions in a NOx/O2 Atmosphere


One of the strategies for lowering soot oxidation temperatures is to
employ a stronger oxidant than oxygen. This is the basis of the so-
called continuously regenerating trap (CRT) developed by Johnson
Matthey.163 These traps contain an upstream Pt oxidation catalyst,
which function as a NO2 generator by oxidation of exhaust NO. Due
to the higher soot oxidation potential of NO2 compared to oxygen,
a significant decrease in the oxidation temperature is observed,
because of the direct reaction of NO2 with carbon. Systems in which
the catalyst promotes the oxidation of NO to NO2 are called ‘indirect
catalysts’; in these cases it is NO2 that directly oxidizes soot and the
type of soot–catalyst contact is less critical than in oxygen-catalysed
soot oxidation.

b1469_Ch-11.indd 600 4/8/2013 12:39:07 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Formulations for Catalysts for Diesel Soot Combustion 601

Table 11.6 Soot oxidation temperatures (Tm) under NOx/O2 atmosphere.


Soot/catalyst
Aging ratio (type of
temperature contact)soot O2(%); Tox
Sample (°C) type NOx (ppm) (°C) Reference
5%Cu,2%K/Ce 500 1:10 (tight) 10%; 1000 313 169
Printex U
Cu–Ce 500 1:10 (tight) 9.5%; 1000 320 170
Printex U
Co/CeO2-cop 400 1:20 (tight) 6%; 5000 320 76
real
CeO2 500 1:10 (tight) 10%; 1000 333 169
Printex U
CeO2 500 1:10 (tight) 9.5%; 1000 336 170
Printex U
Cu–Ce–Zr 500 1:10 (tight) 9.5%; 1000 342 170
Printex U
Ce0.4Zr0.5Pr0.1O4 600 1:10 (tight) 5%; 2000 361 168
Printex U
Ce–Zr 500 1:10 (tight) 9.5%; 1000 367 170
Printex U
12%Co,16% 400 1:20 (tight) 18%; 40000 368 45
Ba,7%K/CeO2 real
Ce0.70Zr0.30O2 600 1:10 (tight) 5%; 2000 374 168
Printex U
Cu–Ce 800 1:10 (tight) 9.6%; 1150 380 171
Printex U
2%Cu/5%K/ 500 1:10 (loose) 10%; 1000 359 169
Ce Printex U
Co20/nmCeO2 600 1:10 (loose) 5%; 2000 368 148
Printex U
Cu–Ce 500 1:10 (loose) 9.5%; 1000 417 170
Printex U
Ce0.90Pr0.10O2−δ 1000 1:4 (loose) 10%; 600 419 65,105
Printex U
(Continued )

b1469_Ch-11.indd 601 4/8/2013 12:39:07 PM


b1469 Catalysis by Ceria and Related Materials

602 E. Aneggi, C. de Leitenburg and A. Trovarelli

Table 11.6 (Continued )


Soot/catalyst
Aging ratio (type of
temperature contact)soot O2(%); Tox
Sample (°C) type NOx (ppm) (°C) Reference
Ce0.90Pr0.10O2−δ 500 1:4 (loose) 10%; 600 440 66
Printex U
CeO2 500 1:4 (loose) 10%; 600 442 66
Printex U
Cu–Ce–Zr 500 1:10 (loose) 9.5%; 1000 448 170
Printex U
Mn25Ce75 650 1:20 (loose) 10%; 1000 450 173
real
CeO2 450 1:2 (loose) 10%; 700 452 165
Printex U
Ce0.90La0.10O2−δ 1000 1:4 (loose) 10%; 600 460 65,105
Printex U
Ce0.90La0.10O2−δ 500 1:4 (loose) 10%; 600 465 66
Printex U
CeO2 500 1:10 (loose) 9.5%; 1000 465 170
Printex U
CeO2 1000 1:4 (loose) 10%; 600 474 65,105
Printex U
Ce–Zr 500 1:10 (loose) 9.5%; 1000 475 170
Printex U
Cu–Ce 800 1:10 (loose) 9.6%; 1150 477 171
Printex U
CeO2 500 1:10 (loose) 10%; 1000 480 169
Printex U
Ce0.90Sm0.10O2−δ 1000 1:4 (loose) 10%; 600 493 65,105
Printex U
Ce0.90Y0.10O2−δ 1000 1:4 (loose) 10%; 600 504 65,105
Printex U
Ce0.50Zr0.50O2 700 1:10 (loose) 10%; 1000 500 67
Printex U

(Continued )

b1469_Ch-11.indd 602 4/8/2013 12:39:07 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Formulations for Catalysts for Diesel Soot Combustion 603

Table 11.6 (Continued )


Soot/catalyst
Aging ratio (type of
temperature contact)soot O2(%); Tox
Sample (°C) type NOx (ppm) (°C) Reference
CeO2 500 1:4 (loose) 5%; 500 515 166,167
Printex U
Ce0.76Zr0.24O2 500 1:4 (loose) 5%; 500 517 166,167
Printex U
CeO2 600 1:4 (loose) 5%; 500 550 164
Printex U
Ce0.56Zr0.44O2 500 1:4 (loose) 5%; 500 587 166,167
Printex U
Ce0.36Zr0.64O2 500 1:4 (loose) 5%; 500 595 166,167
Printex U
Ce0.16Zr0.84O2 500 1:4 (loose) 5%; 500 607 166,167
Printex U

Several investigations have been carried out on the catalytic activ-


ity of ceria and ceria-based materials under a NOx/O2 atmosphere,
both in the presence and in the absence of noble metals148,164–176 (see
Table 11.6 for a summary of the main results). As reported177 the
ability of ceria to oxidize NO to NO2 is well established, and this is
the basis of the enhancement observed when soot oxidation is car-
ried out with a mixture of NO/oxygen compared to oxygen.
As shown in Fig. 11.11, the Tm obtained in the presence of NO
is shifted considerably to lower temperatures with respect to that
obtained in the presence of O2 only. It should also be observed that
pretreatment with NO/O2176 or with NO2/O2165 also facilitates soot
oxidation. A bimodal shape of the TPO profile is observed for the
NO + O2 pretreated ceria with the low and high temperature peaks
correlating well with those obtained in NO + O2 and O2 respectively
(Fig. 11.11). This suggests that NO2 derived from adsorbed nitrates
significantly promotes soot oxidation.
The formation of nitrate and NO derived NO2 is not the only
way to explain in detail the soot oxidation potential of ceria under

b1469_Ch-11.indd 603 4/8/2013 12:39:08 PM


b1469 Catalysis by Ceria and Related Materials

604 E. Aneggi, C. de Leitenburg and A. Trovarelli

5000
b
CO2 concentration (ppm) 4000
a
3000

2000 c

1000

0
300 400 500 600
Temperature (°C)

Figure 11.11 TPO profiles of CeO2, in different atmospheres. a: NO+O2; b: O2;


c: O2 pretreated in NO+O2.176 Reprinted with permission from Elsevier.

a NO/O2 atmosphere. A series of careful experiments performed


by Setiabudi et al.165 using different sequences of reactor configura-
tions showed that the interaction of ceria with NO2 originates
some form of active oxygen, which plays an important role in the
acceleration of soot oxidation. The formation of active oxygen
originates from adsorption of NO2 to form nitrites and nitrates
initially, which on decomposition give NO and ‘stored’ oxygen in
the form of cerium peroxide or superoxide, as in the scheme
below:165

This stored oxygen is then actively used to accelerate soot oxida-


tion. These features characterize the peculiar behaviour of ceria in
comparison to other supports like TiO2 and ZrO2, which are not
able to promote oxidation of NO to NO2 or provide active oxygen
species.164

b1469_Ch-11.indd 604 4/8/2013 12:39:08 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Formulations for Catalysts for Diesel Soot Combustion 605

In short the mechanism of the reactions can be summarized as


follows: CeO2 catalyses the oxidation of NO to NO2, as a result NO2
is available for the NO2–soot reaction. In addition, at low tempera-
tures CeO2 is able to adsorb NO2 in the form of nitrates. These are
then decomposed to NO2, which again contributes to accelerate the
soot–NOx reactions:

NO + Ocat → NO2 (11.4)


NO3− → ½O2 + NO2 (11.5)

The nitrates and NO-derived NO2 contribute to the oxidation of


carbon soot by forming surface oxygen complexes C(O):

Cs + NO2 → C(O) + NO (11.6)

These surface complexes comprise any organic intermediate con-


taining oxygen like carboxylic anhydrides, lactones, ethers, etc.176
These intermediates can also be produced via the adsorbed ‘active
oxygen species’ derived from cerium peroxides or superoxides
formed by NO2 interaction:

Cs + Ocat* → C(O) (11.7)

The decomposition of C(O) species occurs through a reaction with


oxygen or NO2:

C(O) + O2 → CO2 + Cs (11.8)


C(O) + NO2 → NO + CO2 + Cs (11.9)

The doping of ceria with zirconia results in active catalysts and the
activity depends strongly on composition;166,167 a relation between
the amount of ceria and catalytic activity is shown in Fig. 11.12. Pure
ceria and Ce0.76Zr0.24O2 show the highest activity but when the con-
tent of zirconia increases, activity progressively decreases indepen-
dently of the surface area. The increase of activity has been
correlated to the NO2 production capacity of the different catalysts,
which is also affected by the production and adsorption strength of

b1469_Ch-11.indd 605 4/8/2013 12:39:08 PM


b1469 Catalysis by Ceria and Related Materials

606 E. Aneggi, C. de Leitenburg and A. Trovarelli

650 70

60

50

Surface area (m2/g)


600
40
Tm (°C)

30
550
20

10

500 0
0 10 20 30 40 50 60 70 80 90 100
Amount of ceria (mol %)

Figure 11.12 Surface area and activity dependence vs. amount of ceria for ceria–
zirconia solid solutions. „: Tm; c : surface area. Elaboration of data from Atribak
et al.167 Reprinted with permission from Elsevier.

nitrates. It was found that Zr-rich compositions contain strong NOx


adsorption sites requiring higher temperatures to transform nitrites
into nitrates thus explaining their poor activity compared to cerium-
rich formulations. Matyshak et al.172 investigating ceria–zirconia
doped catalysts found a strict correlation between the temperature
of the onset of desorption of nitrates from the supports and the
onset of soot combustion, supporting the direct participation of
nitrate complexes in soot oxidation.
A series of other investigations were carried out on the active
Ce0.76Zr0.24O2 composition. Atribak et al.178 correlated the catalytic
activity of several samples of Ce0.76Zr0.24O2 prepared by different
methods with the surface area and the Ce/Zr atomic surface ratio.
They found that the most active catalysts have high surface area and
a Ce/Zr surface ratio close to the nominal value. Surface area alone
cannot explain the trend of activity and also the homogeneity of
ceria distribution plays a key role. The authors also investigated the
correlation between NO2 production and soot oxidation over the
same composition.179 BET surface area and Ce/Zr atomic surface

b1469_Ch-11.indd 606 4/8/2013 12:39:08 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Formulations for Catalysts for Diesel Soot Combustion 607

ratio influence the nitrite to nitrate transformation, which affects


soot oxidation, confirming the main role of the total amount of NO2
available for particulate combustion. In addition, over the same
catalyst formulation, the authors analysed in detail the uncatalysed
and catalysed combustion of two commercial carbon blacks and
three diesel soot samples.180 Real particulates are more reactive than
synthetic soot due to a higher content of heteroatoms (O, H and N)
and the authors observed that uncatalysed and catalysed reactions
follow the same tendency under a NOx+O2 atmosphere, the latter
being always shifted towards a lower temperature.
The doping of ceria with other rare earths like La and Pr high-
lights the importance of redox properties.65,66 Rare-earth modifica-
tion of ceria results in a lowering of the soot oxidation temperature
of ca. 100°C in the case of Pr and La. NO2 produced by reaction with
lattice oxygen is responsible for the activity and its formation is
directly connected to the improvement of redox properties in the
case of Pr and to an increase in surface area in the case of La. The
modification of ceria with a cation that can be more easily reduced
(Pr3+/4+) provides a more reducible surface leading to the improved
formation of NO2 and consequently to higher soot oxidation activ-
ity. Moreover, in CePrOx the synergism between Ce4+ and Pr4+ leads
to active oxygen (the oxygen held between the Ce4+–O–Pr4+), which
enhances NO oxidation; indeed, increasing the amount of praseo-
dymium in the solid solution increases catalytic activity, which
reaches a maximum for a composition with 50 wt% loading of pra-
seodymium. In contrast, La3+ doping increases the number of oxy-
gen vacancies and consequently decreases the concentration of
surface oxygen, which is needed to convert NO to NO2. In this case,
the increase in surface area was shown to compensate for the lower
redox activity. For the same reason CeSmOx and CeYOx were found
to be less active than pure ceria.
Several other transition-metal oxides have been used in combi-
nation with ceria and ceria–zirconia for the simultaneous removal of
NOx and soot. In Cu-doped ceria and ceria–zirconia, the incorpora-
tion of copper enhanced the NO oxidation ability of the ceria cata-
lyst. The strong synergism between Cu and Ce species improves

b1469_Ch-11.indd 607 4/8/2013 12:39:08 PM


b1469 Catalysis by Ceria and Related Materials

608 E. Aneggi, C. de Leitenburg and A. Trovarelli

oxygen activation and NO adsorption by the catalyst promoting soot


oxidation. Thermal resistance and activity can be enhanced in Cu–
Ce–Al mixed oxides, where Cu–ceria interactions benefit from the
enhanced textural properties of alumina.169–171,181
In recent years MnOx–CeO2 mixed oxides have been investi-
gated as a group of relatively cheap and active catalysts for soot
oxidation in the presence of NOx.176,182–185 The catalytic activity of
these formulations seems to be correlated to the enhanced capac-
ity of oxygen transfer and to NO oxidation and NOx adsorption
features. Soot oxidation is strictly correlated with NO2 production
and is influenced by the surface area, the surface ratio Mn/Ce, the
dispersion of Mn species on CeO2 and the oxidation state of Mn
species. MnOx–CeO2 mixed oxides are characterized by a syner-
getic effect that results in higher specific surface area, improved
low-temperature redox properties and smaller oxide crystallites.
Due to the good starting activity of these materials a few attempts
have been made to enhance their oxidation properties.183 The
impregnation of MnOx–CeO2 in alumina powder increases the
textural stability of the catalyst thus maintaining the strong syner-
getic effect between the two oxides despite the lower oxygen avail-
ability due to the presence of the inert support. The loading of
barium on a MnOx–CeO2 mixed oxide improves the NOx storage
capacity of the mixed oxides and despite the loss of surface area
due to the addition of barium the ternary catalyst shows enhanced
NOx-assisted soot oxidation.182 The presence of barium indeed
enhanced the formation of bidentate and monodentate nitrate
coordinates with oxides that are correlated with soot oxidation
activity.
Cobalt in the form of the oxide Co3O4 or as a solid solution was
reported to decrease the reduction temperatures of ceria by several
hundreds of degrees148,186 and this affects nitrate storage capacity,
which in turn will result in a higher activity for the NO2–soot
reaction.
Cobalt- and copper-containing perovskites such as La0.9K0.1CoO3
and La1.7Rb0.3CuO4 have also recently been deposited on CeO2 nano-
particles for the simultaneous removal of soot and NOx.187,188

b1469_Ch-11.indd 608 4/8/2013 12:39:08 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Formulations for Catalysts for Diesel Soot Combustion 609

The aim was to overcome the low surface area of perovskites using a
support that is ‘inert’ toward their metal component, thus prevent-
ing the formation of inactive components. In addition the presence
of ceria might help to enhance the activity of perovskites. The rea-
sons for the enhancement of activity were related to the augmented
adsorption of oxygen species O− and O2− on oxygen vacancies. This
increase is largely due to the better redox properties of the perovskite–
ceria combination compared to the separate materials.
The enhancement of NOx adsorption through the formation of
nitrate species can be obtained by doping ceria with potassium due
to increased basicity. This effect can be augmented with the addi-
tion of barium to obtain formulations with good activity for the
simultaneous removal of soot and NOx, working either as oxidation
catalysts or NOx storage traps.45–47 In a Ba,K/CeO2 catalyst the bar-
ium is an efficient NOx trap; however, the formation of stable
nitrates species causes a loss of soot combustion activity. When Co is
incorporated in the system, the barium nitrates become unstable
and can be readily decomposed to N2 under a reducing atmosphere,
making this system very interesting as an NOx catalytic trap, which
favours soot combustion. The higher activity of Co,Ba,K/CeO2 is
probably due to the formation of the perovskite BaCoO2.93.45 This
structure is correlated with the formation of N-bound nitrate species
(O–Ba–NO2). In this case the barium is chemically bound to the
cobalt forming a mixed oxide and the formation of Ba(NO3)2 is
inhibited. In addition the Ba,K/CeO2 formulation seems to prevent
sulfation of potassium175 (which was observed in the absence of
NO68), thus preserving the activity of K towards soot combustion.
This occurs through the thermodynamically favoured reaction:
K2SO4 + Ba(NO3)2 → KNO3 + BaSO4.
A synergic effect due to the simultaneous doping of CeO2 by potas-
sium and copper has been recently reported.181 The lowering of oxida-
tion temperatures to ca. 330°C was obtained by an increased NOx
storage capacity brought about by K and Cu as shown in Fig. 11.13.
A strong enhancement of soot oxidation activity was also reported
with K-impregnated ceria–zirconia. The addition of potassium dra-
matically enhanced the catalytic activity of bare Ce0,5Zr0,5O2 due to the

b1469_Ch-11.indd 609 4/8/2013 12:39:08 PM


b1469 Catalysis by Ceria and Related Materials

610 E. Aneggi, C. de Leitenburg and A. Trovarelli

800

d
NO concentration (x10-6) 600

400
b c

200
a

100 200 300 400 500 600 700


Temperature (°C)

Figure 11.13 NOx-TPD curves over a: CeO2; b: K/Ce; c: K/Cu; d: K/Cu/Ce.


Elaboration of data from Weng et al.181 Reprinted with permission from Elsevier.

promotional effect of potassium on the NO storage capacity and the


soot–catalyst contact.189

11.7 Conclusions
Ceria-based formulations are among the more active catalysts for
soot oxidation either under O2 or in a NOx/O2 atmosphere, which
decreases the combustion temperature from above 600°C down to
320–350°C. These results are very important for fulfilling new regu-
lations on pollutant emission control. Among the several physico-
chemical parameters that influence the overall activity, the redox
capacity and the availability of surface ‘active oxygen species’ are
certainly the most important.
There is still much work to do in the direction of improving
activity and stability under different reaction conditions and meet-
ing upcoming technological needs. The more recent strategies for
reducing pollutant emissions from diesel engines are based on a
synergetic combination of existing catalytic after-treatment tech-
nologies in order to obtain performance in the combined system

b1469_Ch-11.indd 610 4/8/2013 12:39:08 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Formulations for Catalysts for Diesel Soot Combustion 611

that exceeds those of each method considered separately. The new


stringent regulation limits on nitrogen oxides (NOx) and soot emis-
sions are forcing the development of innovative catalytic systems
that combine NOx treatment with catalytic filters simplifying and
reducing the size of the exhaust gas after-treatment system. In this
framework a few investigations have reported on ceria-based cata-
lysts that have optimal activity in selective catalytic reduction and
good activity in soot oxidation.190–192 On the material side, new fron-
tiers for further progress would be the synthesis of novel formula-
tions based on a combination of transition or noble metals and on
the other hand the modification of the chemico-physical parameters
through the synthesis of innovative nanoarchitectures, which can
lead to novel properties such as high surface area and controlled
morphology, to achieve higher activity.

Acknowledgments
Financial support from the EU Italy-Austria Interreg Mat4Cata
project and Regione FVG is gratefully acknowledged.

References
1. Summers, J.C., Van Houtte, S., Psaras, D., Appl. Catal. B, 10 (1996)
139–156.
2. Majewski, W.A., Khair, M.K., Diesel Emissions and Their Control, SAE
International, Warrendale, PA, (2006).
3. Neeft, P.A., Makkee, M., Moulijn, J.A., Appl. Catal. B, 8 (1996) 57–78.
4. van Setten, B.A.A.L., Makkee, M., Moulijn, J.A., Catal. Rev., 43 (2001)
489–564.
5. http://www.cdc.gov/NIOSH. Accessed 16 June 2011.
6. http://www.iarc.fr. Accessed 16 June 2011.
7. Siegmann, Q.Z.K., Keller, A., Matter, U., Scherrer, L., Siegmann, H.C.,
Atmos. Environ., 34 (2000) 443–451.
8. http://www.dieselnet.com. Accessed 16 June 2011.
9. Millet, C.N., Chedotal, R., Da Costa, P., Appl. Catal. B, 90 (2009)
339–346.

b1469_Ch-11.indd 611 4/8/2013 12:39:08 PM


b1469 Catalysis by Ceria and Related Materials

612 E. Aneggi, C. de Leitenburg and A. Trovarelli

10. Weibel, M., Waldbusser, N., Wunsch, R., Chatterjee, D., Bandl-Konrad,
B., Krutzsch, B., Top. Catal., 52 (2009) 1702–1708.
11. Twigg, M., Appl. Catal. B, 70 (2007) 2–15.
12. Fino, D., Sci. Techn. Adv. Mater., 8 (2007) 93–100.
13. Zelenka, P., Cartellieri, W., Herzog, P., Appl. Catal. B, 10 (1996) 3–28.
14. Stratakis, G.A., Stamatelos, A.M., Combust. Flame, 132 (2003) 157–169.
15. Shelef, M., McCabe, R.W., Catal. Today, 62 (2000) 35–50.
16. Mul, G., Zhu, W., Kapteijn, F., Moulijn, J.A., Appl. Catal. B, 17 (1998)
205–220.
17. Jelles, S.J., van Setten, B.A.A.L., Makkee, M., Moulijn, J.A., Appl. Catal.
B, 21 (1999) 35–49.
18. Liu, S., Obuchi, A., Uchisawa, J., Nanba, T., Kushiyama, S., Appl. Catal.
B, 37 (2002) 309–319.
19. Oi-Uchisawa, J., Obuchi, A., Enomoto, R., Xu, J., Nanba, T., Liu, S.,
Kushiyama, S., Appl. Catal. B, 32 (2001) 257–268.
20. Peralta, M.A, Gross, M.S., Sanchez, B.S., Querini, C.A., Chem. Eng. J.,
152 (2009) 234–241.
21. Issa, M., Petit, C., Brillard, A., Brilhac, J.-F., Fuel, 87 (2008) 740–750.
22. Saab, E., Abi-Aad, E., Bokova, M.N., Zhilinskaya, E.A., Aboukaïs, A.,
Carbon, 45 (2007) 561–567.
23. Simonsen, S.B., Dahl, S., Johnson, E., Helveg, S., J. Catal., 255 (2008)
1–5.
24. Saab, E., Aouad, S., Abi-Aad, E., Bokova, M.N., Zhilinskaya, E.A.,
Aboukaïs, A., Kinet. Catal., 48 (2007) 841–846.
25. Hensgen, L., Stowe, K., Catal. Today, 159 (2011) 100–107.
26. Stanmore, B.R., Brilhac, J.F., Gilot, P., Carbon, 39 (2001) 2247–2268.
27. Neeft, J.P.A., van Pruissen, O.P., Makkee, M., Moulijn, J.A., Appl. Catal.
B, 12 (1997) 21–31.
28. van Setten, B.A.A.L., Schouten, J.M., Makkee, M., Moulijn, J.A., Appl.
Catal. B, 28 (2000) 253–257.
29. Neeft, J.P.A., Makkee, M., Moulijn, J.A., Chem. Eng. J., 64 (1996)
295–302.
30. Trovarelli, A., ed., Catalysis by Ceria and Related Materials, Imperial
College Press, London, (2002).
31. Bernal, S., Kašpar J., Trovarelli, A., Catal. Today, 50 (1999) 173–443.

b1469_Ch-11.indd 612 4/8/2013 12:39:08 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Formulations for Catalysts for Diesel Soot Combustion 613

32. Trovarelli, A., de Leitenburg, C., Boaro, M., Dolcetti, G., Catal. Today,
50 (1999) 353–367.
33. Kašpar, J., Fornasiero, P., J. Solid State Chem., 171 (2003) 19–29.
34. Kašpar, J., Fornasiero, P., Graziani, M., Catal. Today, 50 (1999)
285–298.
35. Farrauto, R.J., Heck, R.M., Catal Today, 51 (1999) 351–360.
36. Trovarelli, A., de Leitenburg, C., Dolcetti, G., Chem. Tech, June (1997)
32–37.
37. Trovarelli, A., Fornasiero P, eds., Catalysis by Ceria and Related Materials,
2nd ed., Imperial College Press, London (2012).
38. Trovarelli, A., Catal. Rev. Sci. Eng., 38 (1996) 439–520.
39. Bernal, S., Calvino, J.J., Cauqui, M.A., Gatica, J.M., Larese, C., Perez
Omil, J.A., Pintado, J.M., Catal. Today, 50 (1999) 175–206.
40. Trovarelli, A., Boaro, M., Rocchini, E., de Leitenburg, C., Dolcetti, G.,
J. Alloys Compd., 323 (2001) 584–591.
41. Kašpar, J., Graziani, M., Fornasiero, P., in Handbook on the Physics and
Chemistry of Rare Earths, vol. 29, eds. K.A. Gschneidner, Jr., L. Eyring,
Elsevier Science B.V., Amsterdam, (2000), pp. 159–267.
42. Di Monte, R., Kašpar, J., Top. Catal., 28 (2004) 47–58.
43. Saracco, G., Badini, C., Russo, N., Specchia, V., Appl. Catal. B, 21
(1999) 233–242.
44. Ciambelli, P., Palma, V., Russo, P., Vaccaro, S., J. Mol. Catal. A, 204–205
(2003) 673–681.
45. Milt, V.G., Querini, C.A., Mirò, E.E., Ulla, M.A., J. Catal., 220 (2003)
424–432.
46. Pisarello, M.L., Milt, V., Peralta, M.A., Querini, C.A., Mirò, E.E., Catal.
Today, 75 (2002) 465–470.
47. Mirò, E.E., Ravelli, F., Ulla, M.A., Cornaglia, L.M., Querini, C.A.,
Catal. Today, 53 (1999) 631–638.
48. Ciambelli, P., Palma, V., Russo, P., Vaccaro, S., Catal. Today, 73 (2002)
363–370.
49. Badini, C., Saracco, G., Serra, V., Specchia, V., Appl. Catal. B, 18
(1998) 137–150.
50. Neri, G., Rizzo, G., Galvagno, S., Musolino, M.G., Donato, A.,
Pietropaolo, R., Thermochim. Acta, 381 (2002) 165–172.

b1469_Ch-11.indd 613 4/8/2013 12:39:08 PM


b1469 Catalysis by Ceria and Related Materials

614 E. Aneggi, C. de Leitenburg and A. Trovarelli

51. Querini, C.A., Ulla, M.A., Requejo, F., Soria, J., Sedrán, U.A.,
Mirò, E.E., Appl. Catal. B, 15 (1998) 5–19.
52. Jimenez, R., Garcia, X., Cellier, C., Ruiz, P., Gordon, A.L., Appl. Catal.
A, 297 (2006) 125–134.
53. Fino, D., Fino, P., Saracco, G., Specchia, V., Korean J. Chem. Eng., 20
(2003) 445–450.
54. An, H., Kilroy, C., McGinn, P.J., Catal. Today, 98 (2004) 423–429.
55. Russo, N., Furfori, S., Fino, D., Saracco, G., Specchia, V., Appl. Catal.
B, 83 (2008) 85–95.
56. Aneggi, E., Boaro, M., de Leitenburg, C., Dolcetti, G., Trovarelli, A.,
J. Alloys Compd., 408–412 (2006) 1096–1102.
57. Aneggi, E., Boaro, M., de Leitenburg, C., Dolcetti, G., Trovarelli, A.,
Catal. Today, 112 (2006) 94–98.
58. Bueno-López, A., Krishna, K., Makkee, M., Moulijn, J.A., J. Catal., 230
(2005) 237–248.
59. Makkee, M., Jelles, S.J., Moulijn, J.A., in Catalysis by Ceria and Related
Materials, ed. A. Trovarelli, Imperial College Press, London, (2002),
p. 391.
60. Aneggi, E., de Leitenburg, C., Dolcetti, G., Trovarelli, A., Catal. Today,
136 (2008) 3–10.
61. Aneggi, E., de Leitenburg, C., Dolcetti, G., Trovarelli, A., Top. Catal.,
42–43 (2007) 319–322.
62. Aneggi, E., de Leitenburg, C., Dolcetti, G., Trovarelli, A., Catal. Today,
114 (2006) 40–47.
63. Wu, X., Liu, D., Li, K., Li, J., Weng, D., Catal. Commun., 8 (2007)
1274–1278.
64. Krishna, K., Bueno-López, A., Makkee, M., Moulijn, J.A., Appl. Catal.
B, 75 (2007) 189–200.
65. Krishna, K., Bueno-López, A., Makkee, M., Moulijn, J.A., Appl. Catal.
B, 75 (2007) 201–209.
66. Krishna, K., Bueno-López, A., Makkee, M., Moulijn, J.A., Appl. Catal.
B, 75 (2007) 210–220.
67. Zhu, L., Yu, J., Wang, X., J. Hazard. Mater., 140 (2007) 205–210.
68. Peralta, M.A., Milt, V.G., Cornaglia, L.M., Querini, C.A., J. Catal., 242
(2006) 118–130.
69. Farrauto, R.J., Voss, K.E., Appl. Catal. B, 10 (1996) 29–51.

b1469_Ch-11.indd 614 4/8/2013 12:39:08 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Formulations for Catalysts for Diesel Soot Combustion 615

70. Logothrtidis, S., Patsalas, P., Charitidis, C., Mat. Sci. Eng. C, 23 (2003)
803–806.
71. Bokova, M.N., Decarne, C., Abi-Aad, A., Pryakhin, A.N., Lunin, V.V.,
Aboukaïs, A., Thermochim. Acta, 428 (2005) 165–171.
72. Lamonier, J.F., Sergent, N., Matta, J., Aboukaïs, A., J. Therm. Anal.
Calorim., 66 (2001) 645–658.
73. Dernaika, B., Uner, D., Appl. Catal. B, 40 (2003) 219–229.
74. van Doorn, J., Varloud, J., Meriaudeau, P., Perrichon, V., Appl. Catal.
B, 1 (1992) 117–127.
75. Courcot, D., Abi-Aad, E., Capelle, S., Aboukaïs, A., Stud. Surf. Sci.
Catal., 116 (1998) 625–634.
76. Harrison, P.G., Ball, I.K., Daniell, W., Lukinskas, P., Caspedes, M.,
Mirò, E.E., Ulla, M.A., Chem. Eng. J., 95 (2003) 47–55.
77. Primet, M., Garbowski, E., in Catalysis by Ceria and Related Materials, ed.
A. Trovarelli, Imperial College Press, London, (2002), pp. 407–429.
78. Kockrick, E., Schrage, C., Grigas, A., Geiger, D., Kaskel, S., J. Solid State
Chem., 181 (2008) 1614–1620.
79. Palmisano, P., Russo, N., Fino, P., Fino, D., Badini, C., Appl. Catal. B,
69 (2006) 85–92.
80. Aneggi, E., de Leitenburg, C., Trovarelli, A., unpublished results.
81. Issa, M., Petit, C., Mahzoul, H., Aboukaïs, A., Brilhac, J.F., Top. Catal.,
52 (2009) 2063–2069.
82. Issa, M., Mahzoul, H., Brillard, A., Brilhac, J.-F., Chem. Eng. Technol., 32
(2009) 1859–1865.
83. Bassou, B., Guilhaume, N., Lombaert, K., Mirodatos, C., Bianchi, D.,
Energy Fuels, 24 (2010) 4766–4780.
84. Bassou, B., Guilhaume, N., Lombaert, K., Mirodatos, C., Bianchi, D.,
Energy Fuels, 24 (2010) 4781–4792.
85. Iojoiu, E.E., Bassou, B., Guilhaume, N., Farrusseng, D., Desmartin-
Chomel, A., Lombaert, K., Bianchi, D., Mirodatos, C., Catal. Today,
137 (2008) 103–109.
86. Bueno–López, A., Krishna, K., Makkee, M., Moulijn, J.A., Catal. Letters,
99 (2005) 203–205.
87. Machida, M., Murata, Y., Kishikawa, K., Zhang, D., Ikeue, K., Chem.
Mater., 20 (2008) 4489–4494.

b1469_Ch-11.indd 615 4/8/2013 12:39:08 PM


b1469 Catalysis by Ceria and Related Materials

616 E. Aneggi, C. de Leitenburg and A. Trovarelli

88. Gross, M.S., Sánchez, B.S., Querini, C.A., Chem. Eng. J., 168 (2011)
413–419.
89. Li, P., Chen, I.W., Penner-Hahn, J.E., J. Am. Ceram. Soc., 77 (1994)
118–128.
90. Nagai, Y., Yamamoto, T., Tanaka, T., Yoshida, S., Nonaka, T., Okamoto,
T., Suda, A., Sugiura, M., J. Synchrotron Rad., 8 (2001) 616–618.
91. Di Monte, R., Fornasiero, P., Graziani, M., Kašpar, J., J. Alloys Compd.,
275–277 (1998) 877–885.
92. Trovarelli, A., Zamar, F., Llorca, J., de Leitenburg, C., Dolcetti, G.,
Kiss, J.T., J. Catal., 169 (1997) 490–502.
93. Fornasiero, P., Di Monte, R., Rao, J.R., Kašpar, J., Trovarelli, A.,
Graziani, M., J. Catal., 151 (1995) 168–177.
94. Di Monte, R., Kašpar, J., J. Mater. Chem., 15 (2005) 633–648.
95. Trovarelli, A., Comments Inorg. Chem., 20 (1999) 263–284.
96. Shannon, R.D., Prewitt, C.T., Acta Cryst. B, 25 (1969) 925–946.
97. Kašpar, J., Fornasiero, P., Balducci, G., Di Monte, R., Hickey, N.,
Sergo, V., Inorg. Chim. Acta, 349 (2003) 217–226.
98. Yashima, M., in Catalysis by Ceria and Related Materials, 2nd edition, eds.
A. Trovarelli, P. Fornasiero, Imperial College Press, London, (2012).
99. Vlaic, G., Di Monte, R., Fornasiero, P., Fonda, E., Kašpar, J., Graziani, M.,
J. Catal., 182 (1999) 378–389.
100. Vidmar, P., Fornasiero, P., Kašpar, J., Gubitosa, G., Graziani, M.,
J. Catal., 171 (1997) 160–168.
101. He, H., Dai, H.X., Wong, K.W., Au, C.T., Appl. Catal. A, 251 (2003)
61–74.
102. He, H., Dai, H.X., Ng, L.H., Wong, K.W, Au, C.T., J. Catal., 206 (2002)
1–13.
103. Narula, C.K., Haack, L.P., Chun, W., Jen, H.-W., Graham, G.W., J. Phys.
Chem. B, 103 (1999) 3634–3639.
104. Atribak, I., Bueno-López, A., García-García, A., J. Molec. Catal. A, 300
(2009) 103–110.
105. Krishna, K., Bueno-López, A., Makkee, M., Moulijn, J.A., Top. Catal.,
42–43 (2007) 221–228.
106. Fang, P., Luo, M., Lu, J., Cen, S., Yan, X., Wang, X., Thermochim. Acta,
478 (2008) 45–50.
107. Małecka, M.A., Kεpiński, L., Miśta, W., Appl. Catal. B, 74 (2007) 290–298.

b1469_Ch-11.indd 616 4/8/2013 12:39:08 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Formulations for Catalysts for Diesel Soot Combustion 617

108. Atribak, I., Lopez-Suarez, F.E., Bueno-López, A., García-García, A.,


Catal. Today, 176 (2011) 404–408.
109. Atribak, I., Bueno-López, A., García-García, A., Catal. Comm., 9 (2008)
250–255.
110. Zhang, Z., Han, D., Wei, S., Zhang Y., J. Catal., 276 (2010) 16–23.
111. Aneggi, E., de Leitenburg, C., Dolcetti, G., Trovarelli, A., Catal. Today,
181 (2012) 108–115.
112. Liang, Q., Wu, X., Wu, X., Weng, D., Catal. Lett., 119 (2007)
265–270.
113. Reddy, B.M., Bharali, P., Thrimurthulu, G., Saikia, P., Katta, L., Park, S.,
Catal. Lett., 123 (2008) 327–333.
114. Katta, L., Sudarsanam, P., Thrimuirthulu, G., Reddy, B.M., Appl. Catal.
B, 101 (2010) 101–108.
115. Masui, T., Minami, K., Koyabu, K., Imanaka, N., Catal. Today, 117
(2006) 187–192.
116. Imanaka, N., Masui, T., Egawa, T., Imadzu, H., J. Mater. Chem., 19
(2009) 208–210.
117. Liang, Q., Wu, X., Weng, D., Xu, H., Catal. Today, 139 (2008)
113–118.
118. Liang, Q., Wu, X., Weng, D., Lu, Z., Catal. Comm., 9 (2008) 202–206.
119. Zhang, G., Zhao, Z., Liu, J., Jiang, G., Duan, A., Zheng, J., Chen, S.,
Zhou, R., Chem. Commun., 46 (2010) 457–459.
120. Aouad, S., Saab, E., Abi Aad, E., Aboukaïs, A., Catal. Today, 119 (2007)
273–277.
121. Aouad, S., Saab, E., Abi Aad, E., Aboukaïs, A., Kinet. Catal., 48 (2007)
835–840.
122. Sánchez Escribano, V., Fernández López, E., Gallardo-Amores, J.M.,
del Hoyo Martínez, C., Pistarino, C., Panizza, M., Resini, C., Busca, G.,
Combust. Flame, 153 (2008) 97–104.
123. Saab, E., Aouad, S., Abi-Aad, E., Zhilinskaya, E., Aboukaïs, A., Catal.
Today, 119 (2007) 286–290.
124. Aouad, S., Abi-Aad, E., Aboukaïs, A., Appl. Catal. B, 88 (2009)
249–256.
125. Homsi, D., Aouad, S., El Nakat, J., El Khoury, B., Obeid, P., Abi-Aad, E.,
Aboukaïs, A., Catal. Comm., 12 (2011) 776–780.

b1469_Ch-11.indd 617 4/8/2013 12:39:08 PM


b1469 Catalysis by Ceria and Related Materials

618 E. Aneggi, C. de Leitenburg and A. Trovarelli

126. Bueno-López, A., Krishna, K., van der Linden, B., Mul, G., Moulijn, J.A.,
Makkee, M., Catal. Today, 121 (2007) 237–245.
127. Azambre, B., Collura, S., Darcy, P., Trichard, J.M., Da Costa, P.,
García-García, A., Bueno-López, A., Fuel Process Technol., 92 (2011)
363–371.
128. Li, X., Wei, S., Zhang, Z., Zhang, Y., Wang, Z., Su, Q., Gao, X., Catal.
Today, 175 (2011) 112–116.
129. Gross, M.S., Ulla, M.A., Querini, C.A., Appl. Catal. A, 360 (2009) 81–88.
130. Peralta, M.A., Gross, M.S., Ulla, M.A., Querini, C.A., Appl. Catal. A,
367 (2009) 59–69.
131. Cerfontain, M.B., Meijer, R., Kapteijn, F., Moulijn, J.A., J. Catal., 107
(1987) 173–180.
132. Yuh, S.J., Wolf, E.E., Fuel, 62 (1983) 252–255.
133. Milt, V.G., Querini, C.A., Mirò, E.E., Thermochim. Acta, 404 (2003)
177–186.
134. Zhu, Z.H., Lu, G.Q., J. Catal.,197 (1999) 262–274.
135. Zhu, Z.H., Lu, G.Q., Yang, R.T., J. Catal., 192 (2000) 77–87.
136. López-Fonseca, R., Elizundia, U., Landa, I., Gutiérrez-Ortiz, M.A.,
González-Velasco, J.R., Appl. Catal. B, 61 (2005) 150–158.
137. Döbber, D., Kiebling, D., Schmitz, W., Wendt, G., Appl. Catal. B, 52
(2004) 135–143.
138. Terribile, D., Trovarelli, A., de Leitenburg, C., Primavera, A., Dolcetti, G.,
Catal. Today, 47 (1999) 133–140.
139. Shan, W., Ma, N., Yang, J., Dong, X., Liu, C., Wei, L., J. Natural Gas
Chem., 19 (2010) 86–90.
140. Kim, D.J., J. Am. Ceram. Soc., 72 (1989) 1415–1421.
141. Reddy, B.M., Rao, K.N., Catal. Comm., 10 (2009) 1350–1353.
142. Lamonier, J.F., Kulyova, S.P., Lunin, V.V., Zhilinskaya, E.A., Aboukaïs, A.,
J. Therm. Anal. Calorim., 75 (2004) 857–865.
143. Pruvost, C., Lamonier, J.F., Courtcot, D., Abi Aad E., Aboukaïs, A.,
Stud. Surf. Sci. Catal., 130C (2000) 2159–2164.
144. Fu, M., Yue, X., Ye, D., Ouyang, J., Huang, B., Wu, J., Liang, H., Catal.
Today, 153 (2010) 125–132.
145. Muroyama, H., Hano, S., Matsui, T., Eguchi, K., Catal. Today, 153
(2010) 133–135.

b1469_Ch-11.indd 618 4/8/2013 12:39:09 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Formulations for Catalysts for Diesel Soot Combustion 619

146. Lamonier, J.F., Kulyova, S.P., Zhilinskaya, E.A., Kostyuk, B.G., Lunin, W.,
Aboukaïs, A., Kinet. Catal., 45 (2004) 429–435.
147. Dhakad, M., Mitshuhashi, T., Rayalu, S., Doggali, P., Bakardjiva, S.,
Subrt, J., Fino, D., Haneda, H., Labhsetwar, N., Catal. Today, 132
(2008) 188–193.
148. Liu, J., Zhao, Z., Wang, J., Xu, C., Duan, A., Jiang, G., Yang, Q., Appl.
Catal. B, 84 (2008) 185–195.
149. Flouty, R., Abi-Aad, E., Siffert, S., Aboukaïs, A., Appl. Catal. B, 46
(2003) 145–153.
150. Aneggi, E., Llorca, J., de Leitenburg, C., Dolcetti, G., Trovarelli, A.,
Appl. Catal. B, 91 (2009) 489–498.
151. Shimizu, K., Kawachi, H., Satsuma, A., Appl. Catal., 96 (2010) 169–175.
152. Shimizu, K., Kawachi, H., Komai, S., Yoshida, K., Sasaki, Y., Satsuma, A.,
Catal. Today, 175 (2011) 93–99.
153. Lim, C.B., Kusaba, H., Einaga, H., Teraoka, Y., Catal. Today, 175
(2011) 106–111.
154. Preda, G., Pacchioni, G., Catal. Today, 177 (2011) 31–38.
155. Kayama, T., Yamazaki, K., Shinjoh, H., J. Am. Chem. Soc., 132 (2010)
13154–13155.
156. Murrel, L.L., Carlin, R.T., J. Catal., 159 (1996) 479–490.
157. Kundakovic, Lj., Flytzani-Stephanopoulos, M., Appl. Catal. A, 183
(1999) 35–51.
158. Nagy, A., Mestl, G., Appl. Catal. A, 188 (1999) 337–353.
159. Qu, Z., Cheng, M., Huang, W., Bao, X., J. Catal., 229 (2005) 446–458.
160. Gungor, N., Isci, S., Gunister, E., Mista, W., Teterycz, H., Klimkiewicz, R.,
Appl. Clay Sci., 32 (2006) 291–296.
161. Qu, Z., Cheng, M., Dong, X., Bao, X., Catal. Today, 93–95 (2004)
247–255.
162. Waterhouse, G.I.N., Bowmaker, G.A., Metson, J.B., Appl. Surf. Sci., 214
(2003) 36–51.
163. Cooper, B.J., Roth, S.A., Platinum Metals Rev., 35 (1991) 178–187.
164. Atribak, I., Such-Basánez, I., Bueno-López, A., García-García, A.,
J. Catal., 250 (2007) 75–84.
165. Setiabudi, A., Chen, J., Mul, G., Makkee, M., Moulijn, J.A., Appl. Catal.
B, 51 (2004) 9–19.

b1469_Ch-11.indd 619 4/8/2013 12:39:09 PM


b1469 Catalysis by Ceria and Related Materials

620 E. Aneggi, C. de Leitenburg and A. Trovarelli

166. Atribak, I., Bueno-López, A., García-García, A., J. Catal., 259 (2008)
123–132.
167. Atribak, I., Azambre, B., Bueno López, A., García-García, A., Appl.
Catal. B, 92 (2009) 126–137.
168. Liu, J., Zhao, Z., Xu, C., Duan, A., Wang, L., Zhang, S., Catal. Comm.,
8 (2007) 220–224.
169. Weng, D., Li, J., Wu, X., Lin, F., Catal. Comm., 9 (2008) 1898–1901.
170. Wu, X., Liang, Q., Weng, D., Lu, Z., Catal. Comm., 8 (2007) 2110–2114.
171. Wu, X., Lin, F., Weng, D., Li, J., Catal. Comm., 9 (2008) 2428–2432.
172. Matyshak, V.A., Sadykov, V.A., Kuznetsova, T.G., Ukharskii, A.A.,
Khomenko, T.I., Bykhovskii, M.Ya., Sil’chenkova, O.N., Korchak, V.N.,
Kinet. Catal., 47 (2006) 400–411.
173. Tikhomirov, K., Krocher, O., Elsener, M., Wokaun, A., Appl. Catal. B,
64 (2006) 72–78.
174. Tikhomirov, K., Krocher, O., Wokaun, A., Catal. Lett., 109 (2006)
49–53.
175. Milt, V.G., Peralta, M.A., Ulla, M.A., Mirò, E.E., Catal. Comm., 8 (2007)
765–769.
176. Wu, X., Lin, F., Xu, H., Weng, D., Appl. Catal. B, 96 (2010) 101–109.
177. García-García, A., Bueno-López, A., in Catalysis by Ceria and Related
Materials, 2nd edition, eds. A. Trovarelli, P. Fornasiero, Imperial
College Press, London, (2012).
178. Atribak, I., Bueno-López, A., García-García, A.,Top. Catal., 52 (2009)
2088–2091.
179. Atribak, I., Azambre, B., Bueno-López, A., García-García, A., Top.
Catal., 52 (2009) 2092–2096.
180. Atribak, I., Bueno-López, A., García-García, A., Comb. Flame, 157
(2010) 2086–2094.
181. Weng, D., Li, J., Wu, X., Si, Z., J. Rare Earth, 28 (2010) 542–546.
182. Liu, J., Zhao, Z., Lan, J., Xu, C., Duan, A., Jiang, G., Wang, X., He, H.,
J. Phys. Chem. C, 113 (2009) 17114–17123.
183. Liu, J., Zhao, Z., Xu, C., Duan, A., Jiang, G., Ind. Eng. Chem. Res., 49
(2010) 3112–3119.
184. Wu, X., Liu, S., Lin, F., Weng, D., J. Hazard. Mater., 181 (2010) 722–728.
185. Ito, K., Kishikawa, K., Watajima, A., Ikeue, K., Machida, M., Catal.
Comm., 8 (2007) 2176–2180.

b1469_Ch-11.indd 620 4/8/2013 12:39:09 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Formulations for Catalysts for Diesel Soot Combustion 621

186. Liu, J., Zhao, Z., Xu, J., Xu, C., Duan, A., Jiang, G., He, H., Chem.
Comm., 47 (2011) 11119–11121.
187. Wu, X., Liu, S., Weng, D., Lin, F., Ran, R., J. Hazard. Mater., 187 (2011)
283–290.
188. Wu, X., Liu, S., Weng, D., Lin, F., Catal. Comm., 12 (2011) 345–348.
189. Weng, D., Li, J., Weng, D., Si, Z., J. Environm. Sci., 23 (2011) 145–150.
190. Yoshioyuki, M., Yoshichika, H., US Patent 0176969 A1, to DOWA
Electronics Materials Co., Ltd. Tokyo JP, 2011.
191. Collier, J.E., Dotzel, R., Laroze, S.C., Leppelt, R., Millington, P.J.,
Münch, J.W., Rajaram, R.R., Schedel, H., Patent WO 092525 Al to
Johnson Matthey PLC, 2011.
192. Casapu, M., Bernhard, A., Peitz, D., Mehring, M., Elsener, M.,
Kröcher, O., Appl. Catal. B, 103 (2011) 79–84.

b1469_Ch-11.indd 621 4/8/2013 12:39:09 PM


b1469 Catalysis by Ceria and Related Materials

This page intentionally left blank

b1469_Ch-11.indd 622 4/8/2013 12:39:09 PM


b1469 Catalysis by Ceria and Related Materials

CHAPTER 12

CERIA AND ITS USE IN SOLID OXIDE


CELLS AND OXYGEN MEMBRANES
Christodoulos Chatzichristodoulou, Peter T. Blennow,
Martin Søgaard, Peter V. Hendriksen,
and Mogens B. Mogensen
Department of Energy Conversion and Storage,
Technical University of Denmark, Frederiksborgvej 399,
DK-4000 Roskilde, Denmark

12.1 Introduction
The solid oxide fuel cell (SOFC) has been under development over
several decades1 since it was invented by Baur and Preis2 in 1937. In
order to commercialize this high temperature (600–1000 °C) fuel
cell it was necessary to reduce the costs of fabrication and operation.
Ceria-based materials are of potential interest as both pure and
doped ceria may help to decrease the internal electrical resistance
of an SOFC by reducing the ohmic resistance of the electrolyte as
well as the polarization resistance of both the fuel and the air
electrode. Furthermore, the possibility of less fuel pre-treatment
and having a lower water (steam) content in the natural gas fuel
intended for the SOFC, due to the lower susceptibility towards coke
formation of ceria-containing fuel electrodes, may simplify the fuel
cell system. Finally, ceria-based anodes seem less sensitive to poison-
ing by fuel impurities such as sulfur. The same type of cell has been
developed for electrolysis and then called a solid oxide electrolysis
cell (SOEC). As it is basically the same cell in both applications, it is
often referred to as a solid oxide cell (SOC). It is anticipated that

623

b1469_Ch-12.indd 623 4/8/2013 12:40:50 PM


b1469 Catalysis by Ceria and Related Materials

624 C. Chatzichristodoulou et al.

ceria as an electrocatalyst will also be useful in SOEC applications,


but this has not yet been reported in the literature.
Many of the properties of ceria are intimately related to the tol-
erance of the fluorite structure towards high concentrations of
oxide ion vacancies. In order to select the most appropriate doped
ceria composition for a specific application, it is necessary to know
the defect chemistry and thermochemistry of doped ceria and its
relation to the thermomechanical, crystallographic, and transport
properties. A brief description of the defect chemistry of ceria is
provided in Section 12.2 of this chapter, followed by a short presen-
tation of its thermal, crystallographic, and chemical properties and
their dependence on doping and reduction. Thermodynamic data
on x (in CeO2−x) as a function of oxygen partial pressure pO2
and temperature, as well as the effect of substituents and particle
size, will be a central topic of this chapter and will be discussed in
Section 12.3.
The nonstoichiometric mixed ionic and electronic conducting
(MIEC) properties of doped and reduced ceria are best understood
through its defect chemistry. It is possible to change the properties
of doped ceria from a good oxide ion conducting electrolyte with
negligible electronic conductivity into a MIEC with significant elec-
tronic conductivity by changing the dopant (type and concentra-
tion), operating conditions (oxygen partial pressure pO2 and
temperature), and particle size. An overview of the transport prop-
erties of pure and doped ceria is presented in detail in Section 12.4.
Due to its adjustable transport properties, suitably doped ceria
may be used as an electrolyte (which must be an electronic insulator
and a good ion conductor), as discussed in depth in Section 12.5, or
as a dense oxygen separation membrane (which needs to be a
MIEC). A dense oxygen separation membrane must conduct both
O2− and electrons in parallel but opposite directions — a transport
process termed ambipolar oxygen diffusion, even though it is not of
the same nature as ambipolar diffusion in plasma. The use of doped
ceria as a dense oxygen separation membrane is thoroughly
reviewed in Section 12.6. The MIEC properties of ceria are of
interest for SOC fuel and oxygen electrodes as well, because these

b1469_Ch-12.indd 624 4/8/2013 12:40:51 PM


b1469 Catalysis by Ceria and Related Materials

Ceria and its Use in Solid Oxide Cells and Oxygen Membranes 625

properties increase the available area on which the electrode reac-


tions can take place, as described in detail in Section 12.7. The
surfaces of oxygen separation membranes must also possess good
electrocatalytic properties.
The applicability of a given material for a given purpose is natu-
rally also dependent on all the other physical and chemical proper-
ties of the material. A beneficial property of pure and doped ceria is
that it maintains its fluorite structure over a wide range of pO2 and
for all temperatures until its melting point at ca. 2480°C. This is valid
almost irrespective of which kind of cation substitution (doping) is
used. Furthermore, doped ceria is stable in atmospheres containing
reactive gases, which are often present in SOC and oxygen mem-
brane applications, such as CO2 and H2S.
Even though this review is comprehensive, it does not in any way
cover all the relevant literature on ceria. A complete review is simply
not possible due to the enormous amount of available literature.
A search in Scifinder on “cerium oxide, ceria or CeO2” gave 55,250 hits,
“ceria or CeO2 in fuel cells” gave 4,225 hits, and “ceria or CeO2 in mem-
branes” gave 1,057 hits. Therefore, ceria researchers reading this
chapter without finding references to their own papers should not be
discouraged, because most papers are by nature not cited. Papers that
are believed to contain information that forms a “representative picture
of ceria” have been selected without knowing all available literature.

12.2 The Chemistry and Physics of Ceria in Brief


This section gives an introductory description of the chemical and
physical properties of ceria. These properties are closely related to the
fact that ceria is easily reduced to a nonstoichiometric compound,
CeO2−x. Reduced ceria is a MIEC. Further, the high solubility of a large
number of other oxides in the fluorite-structured ceria makes it pos-
sible to gradually change the properties of the material. Ceria with
dissolved foreign oxides is referred to as “doped ceria,” even though a
term like partially substituted ceria would be better, because substitu-
tion of 20–30% of the Ce4+ with e.g. Gd3+ or Ca2+ is often done. It is also
possible to reduce Ce4+ to Ce3+ in partially substituted ceria.

b1469_Ch-12.indd 625 4/8/2013 12:40:52 PM


b1469 Catalysis by Ceria and Related Materials

626 C. Chatzichristodoulou et al.

12.2.1 Defects in pure and doped ceria


When CeO2 is reduced to CeO2−x, defects are created in the form of
Ce3+, which in Kröger–Vink notation3 is written as Ce¢Ce . In early
works2,4–6, it was discussed whether these substitutional negative
••
defects were charge balanced by Ce interstitials, such as Ce•i , Cei ,
••• •••• ∑∑
Cei , and Cei or by oxide ion vacancies, VO . Oxygen self-diffu-
sion measurements by Steele and Floyd7 provided evidence for the
∑∑
predominance of VO . In situ X-ray and neutron diffraction studies
by Faber et al.8,9 also concluded oxide ion vacancies as being the
predominant charge compensating defect in CeO2−x, the number of
interstitial Ce ions being less than 0.1% of the total defect concen-
tration in CeO1.91. The formation of oxide ion vacancies on reduc-
tion of ceria may be written as:

Oxo + 2CexCe ´ VO∑∑ + 2CeCe


¢ + 1 2 O2(g) (12.1a)

or, depending on the gas composition, e.g. as:

Oxo + 2CexCe + H2(g) ´ VO∑∑ + 2CeCe


¢ + H2O(g) (12.1b)

Oxide ion vacancies may also be introduced by doping with oxides


of metal ions with lower valences, e.g. by dissolution of CaO or
Gd2O3:
CeO
CaO æææ
2
¢¢ + VO∑∑ + Oxo
Æ Ca Ce (12.2)
CeO
Gd2O3 æææ
2
¢ + VO∑∑ + 3Oxo
Æ 2GdCe (12.3)

Existing oxide ion vacancies may be annihilated by doping with


oxides of metal ions with higher valences, e.g. by dissolution of
Nb2O5:
CeO
Nb2O5 + VO∑∑ æææ
2 ∑
Æ 2NbCe + 5Oxo (12.4)

Reaction (12.1) may also take place in doped ceria, and doping
will affect the reducibility of ceria, both due to changing the

b1469_Ch-12.indd 626 4/8/2013 12:40:52 PM


b1469 Catalysis by Ceria and Related Materials

Ceria and its Use in Solid Oxide Cells and Oxygen Membranes 627

concentration of VO∑∑ and Ce′Ce , but also due to changing the partial
molar enthalpy of reduction as well as the vibrational frequency of
the lattice ions (or nonconfigurational entropy).

12.2.2 Lattice parameters of pure, doped and reduced ceria


This section deals with the relations between the lattice parameter,
temperature, substituent cation radius, solubility limit of a substitu-
ent oxide, and x (degree of reduction).

12.2.2.1 Thermal expansion


The lattice parameter at room temperature is 0.541134 nm.10 There
is reasonable agreement in the results of the measured thermal
expansion of pure stoichiometric CeO2 up to 1000 °C,11–16 and the
conclusion is that the linear thermal expansion coefficient (TEC)
for pure unreduced ceria is:

At room temperature: (11.0 ± 0.5) 10−6 K−1


Room temperature to 500 °C: (11.5 ± 0.5) 10−6 K−1
Room temperature to 1000 °C: (12.1 ± 0.5) 10−6 K−1

Doped or reduced ceria will have a different TEC depending on


the ionic radius and charge (oxidation state) of the dopant cation.
If the dopant has a larger radius and lower charge than Ce4+ the
TEC will be increased and vice versa. Examples are: 20% Pr4+, TEC =
11.0 × 10−6 K−1,17 20% Pr3+, TEC = 16.0 × 10−6 K−1,17 20% La3+, TEC =
15.0 × 10−6 K−1,18 and 20% Bi3+, TEC = 16.4 × 10−6 K−1.19

12.2.2.2 Solubility of oxides and change in lattice parameter


The lattice parameter of a simple solid solution usually follows
Vegard’s rule (i.e. a linear lattice parameter dependence on the
solute concentration) with good approximation, often for quite
a few percent of foreign cations, depending on circumstances.
Recently, deviations from Vegard’s rule have been described in

b1469_Ch-12.indd 627 4/8/2013 12:40:52 PM


b1469 Catalysis by Ceria and Related Materials

628 C. Chatzichristodoulou et al.

detail for some fluorite-structured oxides.20 It seems that Vegard’s


rule is followed with good approximation up to 15–25% substitution
of Ce4+ with cations that are not too different from Ce4+ in radius.
The slope of this straight line is termed Vegard’s slope. Kim21
published an empirical relation between the ionic radius and charge
of a cation (dopant) dissolved in ceria (and other fluorite-structured
oxides) and the lattice parameter of the fluorite solid solution
phase:
a = 0.5413 + Â(0.0220 Drk + 0.00015Dzk )mk (12.5)

where a (in nanometers) is the lattice constant of the ceria solid


solution at room temperature, ∆rk (in nanometers) is the difference
in ionic radius (rk − rCe) of the kth dopant and the Ce4+ radius, which
in eightfold coordination is 0.097 nm according to Shannon and
Prewitt,22 ∆zk is the valence difference (zk − 4), and mk is the mole
percent of the kth dopant in the form of MOx. Kim’s lattice constant
for pure ceria is 0.000166 nm higher than the now accepted value of
0.541134 nm. Kim derived his relations from a considerable amount
of data, and his analysis of the uncertainty gave a standard error of
3 × 10−4 nm. Hong and Virkar23 analyzed the relation between the
concentration of LnO1.5 in ceria (and zirconia) and the lattice
parameter of the fluorite lattice. Reducing their formula for the sin-
gle doped ceria case (using the same symbols as above) one obtains:

a = 0.54112 nm + 0.023027(rk - 0.1024 nm)m k (12.6)

Kim argued that the solubility limit of a solute depends on the


elastic energy W, which is introduced in the lattice due to differ-
ences in ionic radius. The larger the elastic energy per substituted
ion, the lower the solubility. The relation between W and the
change in lattice parameter ∆a due to formation of a substitutional
solid solution is given by:

W = 6Ga0(∆a)2 (12.7)
where G is the shear modulus of ceria. ∆a is governed by Vegard’s
slope for the given solute. This implies that the highest solubilities

b1469_Ch-12.indd 628 4/8/2013 12:40:52 PM


b1469 Catalysis by Ceria and Related Materials

Ceria and its Use in Solid Oxide Cells and Oxygen Membranes 629

should be achieved for oxides with cations that have a radius that
results in Vegard’s slope Sv = 0. This radius is the matching radius rm.
From Eq. (12.5) the following expression can be derived for
Vegard’s slope in units of nm.mol%−1 for trivalent dopants:

SV 3,K (nm⋅mol%−1) = 0.022 (rk − 0.097 nm) − 1.5 × 10−4 (12.8)

and for divalent dopants:

SV 2,K (nm⋅mol%−1) = 0.022 (rk − 0.097 nm) − 3.0 × 10−4 (12.9)

Another expression may be obtained for trivalent dopants from Eq.


(12.6):

SV 3,H (nm⋅mol%−1) = 0.023027 (rk − 0.1024 nm) (12.10)

The different formulas give different values for Vegard’s slope and
also imply different values of the matching radius rm. Kim’s expres-
sion implies that rm = 0.1106 nm for divalent ions and 0.1038 nm for
trivalent ions, whereas Hong and Virkar’s expression gives rm =
0.1024 nm for trivalent ions. As mentioned elsewhere24 a value of
rm = 0.101 nm may be derived from other sources.
The experimental literature shows that the fluorite structure of
ceria is in fact very tolerant to dissolution of lower valent metal
oxides with cation radii close to rm, to concentrations that in some
cases are above 40 at% at 1500 °C.25,26 However, the solubility limit is
temperature dependent, i.e. the solubility limit is expected to be
lower at lower temperatures. Although the solubility limits of rare
earth elements in ceria at temperatures below 1000 °C is still
unknown, to the best of our knowledge, recent findings indicate
solubility limits of about 20 mol% at 1000 °C. This is discussed in
more detail in Section 12.4. Oxides with cations much smaller or
bigger than rm are less soluble, but there is contradictory informa-
tion in the literature for the solubility limits.27–29 Fair predictions
of solubility limits are possible using Vegard’s slope criteria. For
example, the solubility of CoO and NiO in ceria were estimated to
be below ca. 1 at%,30 in good agreement with later experimental

b1469_Ch-12.indd 629 4/8/2013 12:40:52 PM


b1469 Catalysis by Ceria and Related Materials

630 C. Chatzichristodoulou et al.

results.31 A more advanced thermodynamic tool by which the


solubility limits may be quantitatively estimated exists.32

12.2.2.3 Expansion on reduction: stoichiometric expansion


As the Ce3+ ion is bigger than rm, ceria will expand on reduction. If
the value rCe3+ = 0.1143 nm is inserted into Eq. (12.5), then:

a = 0.5413 nm + mCeO1.5 × 2.306 × 10−4 nm⋅mol%−1 CeO1.5 (12.11)

i.e. a Vegard’s slope of 2.306 × 10−4 nm⋅mol%−1 CeO1.5. Inserting rCe3+


into Eq. (12.6) gives:

a = 0.54112 nm + mCeO1.5 × 2.740 × 10−4 nm.mol%−1 CeO1.5(12.12)

i.e. a Vegard’s slope of 2.740 × 10−4 nm.mol%−1 CeO1.5. As mCeO1.5


(mol%) = 200 × x (in CeO2−x), we get for homogeneous samples at
room temperature using (12.12):

a = 0.54112 nm + x × 5.48 × 10−2 nm (12.13)

Figure 12.1 shows a plot of the percentage expansion versus the


oxygen stoichiometry N = 2−x at 900 °C. Using the lattice parameter
of 0.54665 nm at 900 °C and the slope of the full line in Fig. 12.1 we
find a Vegard’s slope of 2.71 × 10−4 nm⋅mol%−1 CeO1.5. This is in
good agreement with the radius relations, i.e. with Eq. (12.11) and
in particular with Eq. (12.12). Thus the volume change due to
expansion on reduction is directly proportional to the change in
average ionic radius in the system or the change in oxygen content.
This is often expressed by the linear strain ε:

DL
e= = Sx
L (12.14)

where L is the sample length, ∆L is the change in length, S is the


stoichiometric expansion coefficient, and x is the oxygen nonstoi-
chiometry in CeO2−x. The stoichiometric expansion coefficient is
also called the chemical expansion coefficient. This phenomenon is

b1469_Ch-12.indd 630 4/8/2013 12:40:52 PM


b1469 Catalysis by Ceria and Related Materials

Ceria and its Use in Solid Oxide Cells and Oxygen Membranes 631

Figure 12.1 Percentage expansion of undoped ceria versus N = 2−x at 900 °C.24
The main data are from Chiang et al.,15 … are from Mogensen,33 and the straight
line is the best linear fit.24 Reprinted with permission from Elsevier.

also known for oxides with other structures (e.g. perovskites), where
it also follows a relation similar to Eq. (12.14).34,35 The expansion
related to changes in oxygen stoichiometry can be distinguished
from the effect of a temperature change on the lattice by carrying
out isothermal dilatometry, where sample dimension changes are
measured as a function of oxygen activity.16,36 Table 12.1 gives exam-
ples of S values for a variety of ceria compositions. Results for pure
ceria are included for comparison.

12.3 Defect Chemistry and Thermodynamic Properties


of Pure and Doped Ceria
12.3.1 Defects in pure and doped ceria
The quasi-chemical reactions (12.1)–(12.4), introduced in Section
12.2, form the basis for the description of the defect chemistry and

b1469_Ch-12.indd 631 4/8/2013 12:40:52 PM


b1469 Catalysis by Ceria and Related Materials

632 C. Chatzichristodoulou et al.

Table 12.1 Expansion on reduction of pure and doped ceria.17,24 S


is the stoichiometric expansion coefficient. The estimated uncer-
tainty on S is in the range of 0.01.
Cation % Dopant x % expansion S (mol−1)
0 0.105 1.14 0.11
0 0.200 2.10 0.11
3+
18 Gd 0.060 0.77 0.13
3+
18 Gd 0.120 1.49 0.12
3+
40 Gd 0.022 0.28 0.13
3+
40 Gd 0.045 0.67 0.15
20 Pr → Pr
4+ 3+a
0.097 0.82 0.08
3+
20 Pr 0.138 1.47 0.11
20 Tb → Tb
4+ 3+a
0.089 0.80 0.09
3+
20 Tb 0.153 1.40 0.09
a
Note: Expansion related to reduction of the dopant.

thermochemistry of pure and doped ceria. In addition, protons and


electron holes may be created in ceria by the reaction of water vapor
or oxygen gas, respectively, with oxide ion vacancies:

H2O(g) + VO∑∑ ´ OxO + 2H∑ (12.15)

1
2 O2(g) + VO∑∑ ´ OxO + 2h∑ (12.16)

Applying the law of mass action to Eq. (12.1a) gives:

2 1
È VO∑∑ ˘ ÈCeCe
¢ ˘˚ PO2
KR = Î ˚ Î 2
2
(12.17)
ÈOO ˘ ÈCeCe ˘
x x
Î ˚Î ˚

where square brackets denote concentrations of the respective spe-


cies. In pure ceria, and for small deviations from stoichiometry
¢ ] = 2x , [OxO ] = 2 - x ª 2, and[CeCe ] = 1 - 2x ª 1,
x
(x < 0.01): [VO∑∑ ] = x , [CeCe

b1469_Ch-12.indd 632 4/8/2013 12:40:52 PM


b1469 Catalysis by Ceria and Related Materials

Ceria and its Use in Solid Oxide Cells and Oxygen Membranes 633

¢ ] = 2[VO∑∑ ], has been


where the principle of charge neutrality, [CeCe
used. Insertion into Eq. (12.17) and rearrangement gives:
- 16
x µ PO (12.18)
2

For heavily doped ceria, and for small deviations from stoichiome-
try, the concentration of oxide ion vacancies is effectively constant,
″ ] for divalent or 2[V ∑∑ ] = [Gd ¢ ] for trivalent dopants,
[VO∑∑ ] = [Ca Ce O Ce
and thus Eq. (12.17) gives:

- 14
x µ PO (12.19)
2

The pO2 exponent in Eqs. (12.18) and (12.19) is characteristic of


the prevailing defect reactions and is often used to distinguish
between defect regimes. In the general case of ceria doped with a
divalent cation (e.g. Ce1–yCayO2–x–y), the pO2 exponent varies with
nonstoichiometry and doping level according to the relation:37

-1
∂ log x Ê 2x 2x 8x ˆ Ê ∂ log K R ˆ
= Á -4 - - - Á1 - 2 ˜ (12.20)
∂ log PO2 Ë x + y 2 - x - y 1 - 2x - y ˜¯ Ë ∂ log PO2 ¯

A similar expression can be deduced for the case of ceria doped with
a trivalent cation (e.g. Ce1–yGdyO2–x–y/2), by simply replacing y with
y/2 in the denominator of the first two fractions on the right-hand
side of Eq. (12.20). Figure 12.2 shows the dependence of the pO2
exponent on nonstoichiometry, as calculated from Eq. (12.20) for
various doping levels. As can be seen from Fig. 12.2, Eqs. (12.18)
and (12.19) are only valid for small values of x (x < 0.01). As the
value of x increases, the pO2 exponent is expected to decrease con-
tinuously (in absolute value), approaching a value of −1/10 as the
value of x approaches 0.25. Furthermore, for intermediate acceptor
doping levels of ca. 1000 ppm, the pO2 exponent is expected to shift
continuously from the −1/4 value of acceptor-doped ceria to
the −1/6 value of pure ceria at small oxygen nonstoichiometry
values (x < 0.01).

b1469_Ch-12.indd 633 4/8/2013 12:40:53 PM


b1469 Catalysis by Ceria and Related Materials

634 C. Chatzichristodoulou et al.

-1/10

-1/6
PO2 exponent

-1/4

pure CeO2
-1/3
10 ppm Ca
1000 ppm Ca
10 mol% Ca
pure CeO2 KR corrected
pure CeO2 experimental

-1/2
-4 -3.5 -3 -2.5 -2 -1.5 -1 -0.5
logx

Figure 12.2 Dependence of pO2 exponent on composition and doping level.

The various lines plotted in Fig. 12.2 are calculated values from
Eq. (12.20), assuming that KR is independent of x or pO2. As dis-
cussed in detail later, this is not valid for ceria. The expected varia-
tion of the pO2 exponent with nonstoichiometry (at 1000 °C),
taking into account the pO2 dependence of KR (calculated from
data by Bevan and Kordis38), is illustrated in Fig. 12.2 using open
circles (pure CeO2 KR corrected). As can be seen, the pO2 depend-
ence of KR, arising from the nonstoichiometry dependence of the
equilibrium Gibbs energy change of reaction (12.1a), drastically
influences the expected variation of the pO2 exponent with oxygen
nonstoichiometry, x. The pO2 exponent is expected to pass through
a maximum (absolute) value of ca. −1/2 at intermediate x values, x
≈ 0.05. This is related to the fact that the equilibrium Gibbs energy
change of reaction (12.1a) passes through a minimum at a similar
oxygen nonstoichiometry value, x ≈ 0.1. The nonstoichiometry
dependence of the pO2 exponent, determined directly from experi-
mental data for x = f(pO2) for 99.99% (nominally) pure ceria by

b1469_Ch-12.indd 634 4/8/2013 12:40:53 PM


b1469 Catalysis by Ceria and Related Materials

Ceria and its Use in Solid Oxide Cells and Oxygen Membranes 635

Panlener et al.,39 is also shown in Fig. 12.2 using open squares (pure
CeO2 experimental). As can be seen, it matches well the expected
behavior of the KR corrected pure ceria for x > 0.001. For x < 0.001,
the pO2 exponent increases (in absolute value) towards −1/4 as x
decreases, a behavior that is characteristic of the presence of accep-
tor impurities.
For nonstoichiometries larger than ∼0.01, interactions between
the defects become significant.40,41 This leads to defect association
and ordering. In pure ceria, the singly charged “dimer” and neutral
“trimer” are anticipated to form according to the defect reactions:

( ) (12.21)

¢ + VO∑∑ ´ CeCe
CeCe ¢ VO∑∑

( ) ( )
∑ x
CeCe ¢ VO∑∑
¢ + CeCe ¢ VO∑∑CeCe
´ CeCe ¢ (12.22)

Similar dimers and trimers may form in doped ceria, but in this case
the lower valent dopants may replace Ce¢Ce in the above quasi-
chemical reactions.42,43 The presence of oxide ion vacancies in mul-
tiple states of ionization (VO∑∑ , VO∑ , VO
x
) has also been proposed in the
4,44–47
past, the respective defect reactions being:

VO∑∑ + CeCe
¢ ´ VO∑ + CexCe (12.23)

VO∑ + CeCe
¢ ´ VOx + CexCe (12.24)

Some researchers have also postulated an association between


Ce¢Ce and singly charged oxide ion vacancies.47
For large deviations from stoichiometry, a variety of discrete
compounds have been reported to exist between CeO2 and Ce2O3,
formed as a result of ordering of the oxide ion vacancies. Bevan48
was the first to provide detailed evidence of a sequence of intermedi-
ate rhombohedral phases. Recent studies49–52 have shown that a
series of commensurate superstructures of the fluorite-type lattice

b1469_Ch-12.indd 635 4/8/2013 12:40:53 PM


b1469 Catalysis by Ceria and Related Materials

636 C. Chatzichristodoulou et al.

form in the Ce–O system with compositions following the homolo-


gous series CenO2n−2m, where n and m are integers (n > m). A thermo-
dynamic model of the phase diagram of Ce–O in the entire range of
compositions, including the formation of CenO2n−2m phases, was
recently presented by Zinkevich et al.53 Part of the calculated Ce–O
phase diagram, between CeO1.702 and CeO2, is shown in Fig. 12.3. As
can be seen, it compares well with experimental results.
In this chapter we focus on the oxygen-deficient fluorite-type
CeO2–x phase (pure and doped), for which oxygen dissociation
measurements by Bevan and Kordis38 have indicated a single-phase
region from CeO2 to CeO1.72 above 685 °C (the F phase in Fig. 12.3)
and the appearance of a miscibility gap below 685 °C (the F’ and F’’
phases in Fig. 12.3). Brauer and Gingerich11 also showed, using
high-temperature X-ray diffraction, the presence of a miscibility gap
that closes at 685 °C for the composition CeO1.92.

12.3.2 Defect thermodynamic properties of pure and doped ceria


12.3.2.1 Pure ceria
The temperature and pO2 dependence of x in CeO2–x has been
reported in a number of publications.11,38,39,44,54–61 Three comple-
mentary sets of data from Bevan and Kordis,38 Panlener et al.,39 and
eq
Riess et al.60 are plotted together in Fig. 12.4 in the form of PO2 - x
isotherms at various temperatures. The data from Bevan and
Kordis38 extend to x > 0.28 where ordered phases appear. Fairly
good consistency exists between the reported data.
The partial molar free energy of formation of CeO2–x of a given
nonstoichiometry, DG O ,x is:
2

eq
DG O2 ,x = RT ln PO2 (12.25)

where POeq is the oxygen partial pressure in equilibrium with CeO2–x


2
at temperature T. The nomenclature DG O ,x is adopted (rather
2
than the commonly used DG O2 ) in order to show that the partial
molar free energy of oxidation of CeO2–x is a function of the

b1469_Ch-12.indd 636 4/8/2013 12:40:53 PM


b1469 Catalysis by Ceria and Related Materials

Ceria and its Use in Solid Oxide Cells and Oxygen Membranes 637

Figure 12.3 Part of Ce–O phase diagram between CeO1.702 and CeO2 as a function
of the mole fraction of oxygen (O/Ce+O). Adapted from Zinkevich et al.53 with
permission from Elsevier.

composition x. Combining this with the relation:


DG O2 ,x = DH O2 ,x - T DSO2 ,x , allows the estimation of the partial
molar enthalpy and entropy of formation from a constant composi-
eq
tion PO2 - T data, using the relations:

DH O2 ,x = R
(
d ln PO2
eq
) (12.26)
d 1T( ) x

DS O2 ,x = -R
(
d T ln PO2
eq
) (12.27)
d (T )
x

DH O2 ,x and DSO ,x are plotted as a function of composition in


2
Fig. 12.5 and Fig. 12.6, respectively, using data from Bevan and
Kordis,38 Campserveux and Gerdanian,44 Panhans and Blumenthal,61
and Panlener et al.39 Good agreement is observed between the

b1469_Ch-12.indd 637 4/8/2013 12:40:53 PM


b1469 Catalysis by Ceria and Related Materials

638 C. Chatzichristodoulou et al.

eq
Figure 12.4 PO2 - x isotherms for CeO2–x at various temperatures. Data from
Bevan and Kordis,38 Panlener et al.,39 and Riess et al.60 Reprinted with permission
from Elsevier.

different datasets. All investigations show a strong composition


dependence of the partial molar quantities. Both DH O2 ,x and
DSO2 ,x show a sigmoidal variation with x in the range −2.8 < log
x < −0.6. Bevan and Kordis concluded that DH O2 ,x and DSO2 ,x are
temperature independent in the temperature range 872–1169 °C.38
The strong x dependence of DH O2 ,x for x > 0.01 indicates that there
is strong interaction among defects at these larger departures from
stoichiometry.
Numerous works have attempted to account for the complex
thermodynamic dependence of the CeO2–x nonstoichiometry on
temperature and oxygen partial pressure. In the late 1960s Kofstad
and Hed proposed a model based on singly and doubly ionized Ce
interstitials.5 Blumenthal et al.6 interpreted the defect structure of

b1469_Ch-12.indd 638 4/8/2013 12:40:54 PM


b1469 Catalysis by Ceria and Related Materials

Ceria and its Use in Solid Oxide Cells and Oxygen Membranes 639

Figure 12.5 Composition dependence of the partial molar enthalpy of CeO2–x


according to Bevan and Kordis,38 Campserveux and Gerdanian,44 Panhans and
Blumenthal,61 and Panlener et al.39 Reprinted with permission from Elsevier.

Figure 12.6 Composition dependence of the partial molar entropy of CeO2–x


according to Bevan and Kordis38 and Panlener et al.39 Reprinted with permission
from Elsevier.

b1469_Ch-12.indd 639 4/8/2013 12:40:54 PM


b1469 Catalysis by Ceria and Related Materials

640 C. Chatzichristodoulou et al.

nonstoichiometric ceria on the basis of a model consisting of triply


and quadruply ionized cerium interstitials and localized electrons.
After the predominance of oxide ion vacancies in CeO2−x was
revealed,7–9 Panlener et al.39 showed that in the composition
range −3.0 < log x < −2.2, DSO2 ,x is consistent with a random distri-
bution of doubly ionized oxygen vacancies, while DH O2 ,x exhibits a
slight dependence on composition. They argued that such behavior
may result from a defect/host lattice interaction rather than a
defect/defect interaction.
A puzzling pO2 exponent of −1/5 was observed by a number of
researchers for small values of x (x ≈ 10−3), in the temperature range
650–1500 °C, by both thermogravimetric and electrical conductivity
measurements.6,39,62 This was originally explained by either (i) a sin-
gle state of ionization of the oxygen vacancies and a composition-
dependent electronic mobility and partial molar enthalpy of
oxygen39 or by (ii) multiple states of oxygen vacancy ionization and
a composition-independent electronic mobility and partial molar
enthalpy of oxygen.45,46,63 Kofstad64 was the first to point out the
important role that impurities may play in the range of small x. Later
Dawicke and Blumenthal37 showed that a pO2 exponent varying con-
tinuously from −1/4 to −1/6 with increasing oxygen nonstoichiom-
etry should indeed be expected from the presence of lower valent
impurities dominating the concentration of oxygen vacancies at
high pO2, as shown in Fig. 12.2. Riess et al.60 measured the chemical
potential of oxygen in nonstoichiometric ceria CeO2–x for 10−4 ≤ x
≤ 10−2 and 823 K ≤ T ≤ 1073 K. They found a pO2 exponent of −1/4
for x ≤ 10−3, in line with the expected behavior when the concentra-
tion of doubly ionized oxygen vacancies is dominated by acceptor-
type impurities (900 ppm of Ca in their case). For 10−3 ≤ x ≤ 10−2 they
determined a pO2 exponent of −1/6, which is characteristic of the
case where the concentration of doubly ionized oxygen vacancies is
determined by reduction of Ce.
Riess et al.60 calculated a partial molar enthalpy of oxygen
DH O2 ,x of −9 eV for 3 × 10−3 ≤ x ≤ 10−2 and a lower value (absolute
value) of −7 eV for x = 10−4. Dawicke and Blumenthal37 estimated a
partial molar enthalpy of oxygen of −9.85 ± 0.06 eV, which was

b1469_Ch-12.indd 640 4/8/2013 12:40:55 PM


b1469 Catalysis by Ceria and Related Materials

Ceria and its Use in Solid Oxide Cells and Oxygen Membranes 641

temperature and composition independent in the temperature


range 1073 K < T < 1473 K and in the oxygen nonstoichiometry
range −3.5 ≤ log x ≤ −2.4, for CeO2−x containing 300 ppm Ca impuri-
ties. Panhans and Blumenthal61 determined the composition
dependence of the partial molar enthalpy of oxygen DH O2 ,x by
constant composition coulometric titration measurements for log x
< −2.4. They obtained a maximum (absolute) value of DH Omax 2 ,x
= −10
eV at log x = −2.79, in agreement with the value of −9.96 eV at log
x = −2.8 determined thermogravimetrically by Panlener et al.39
and the value of −9.85 eV determined by oxygen association pres-
sure measurements by Dawicke and Blumenthal.37 The value of
DH Omax
2 ,x
= −10 eV for 10−3 ≤ x ≤ 10−2 is also supported by the findings
of Bevan and Kordis38 and Campserveux and Gerdanian.44
A pronounced decrease in the value of DH O2 ,x for log x < −2.8,
observed by Panhans and Blumenthal61 and shown in Fig. 12.5, is in
line with the findings of Riess et al.,60 who also used coulometric
titration to access the region of small x. The observed drastic
decrease in DH O2 ,x for log x < −2.8, which was not expected in this
dilute defect range, is due to the onset of hole formation according
to Eq. (12.16). The partial molar enthalpy of this reaction was esti-

mated to be ∆H h = 0.15 eV.61 Combining this value with the maxi-
mum value of DH O2 ,x of DH Omax 2 ,x
= −10 eV, Panhans and Blumenthal
calculated a band gap energy of Eg = 2.58 eV, from the relation:
2E g = DH h - DH Omax2 ,x
/2 . This is in fair agreement with calculated65
and experimental findings66,67 for the energy difference between the
O2p valence band and the localized Ce4f states, indicating a value
of ca. 3 eV, when also taking into consideration the expected
decrease in band gap for increasing temperature.68
For oxygen nonstoichiometries larger than 0.01, though, a
model based on randomly distributed doubly ionized oxide ion
vacancies fails to account for the x(T, PO2), DH O2 ,x (x ) , and
DSO2 ,x (x ) dependence observed in CeO2–x. The decrease in the
absolute value of DH O2 ,x for log x > −2.6 is generally attributed to
defect interactions. A number of attempts to account for this behav-
ior were based on the introduction of additional defect reactions
leading to the formation of a variety of defect associates. Tuller and

b1469_Ch-12.indd 641 4/8/2013 12:40:55 PM


b1469 Catalysis by Ceria and Related Materials

642 C. Chatzichristodoulou et al.

Nowick46 implemented a model based on oxide ion vacancies in dif-


ferent states of ionization, which was successful in representing the
experimental data for T > 1100 °C and x < 0.1. Ling47 developed a
statistical thermodynamic model, including contributions from
short-range and long-range coulombic interactions and generalized
∑∑
exclusion effects, assuming a number of point defects, CeCe ¢ , VO ,
∑ ∑ x
∑∑ ∑
¢ VO ) . The inclusion of
¢ VO ) , (CeCe
VO , , and defect associates, (CeCe
coulombic interactions allowed for composition-dependent partial
molar enthalpies, without the need of additional phenomenological
parameters. The fit of the model to a collection of experimental
data, though, was not very successful at temperatures below 1000 °C
and oxygen nonstoichiometries above 0.01. Nakamura69 put forward
a statistical thermodynamic formulation, based on the local defect/
defect interaction model in the mutual first nearest neighbor coor-
dination sphere. Though based on some idealized and phenomeno-
logical assumptions, he was able to reproduce the strong sigmoidal
variation of both DH O2 ,x and DSO2 ,x as a result of the coupled local
defect/defect configurational changes, using the virtual Gibbs for-
mation energies of the local configurations as parameters. Recently,
Duncan et al.70 and Bishop et al.43 have attempted to account for the
PO2 − x dependence of CeO2–x at 800 °C by introducing the dimer
and trimer defect associates described by Eqs. (12.21) and (12.22).
They achieved reasonable agreement with experimental data for
x < 0.1, but the model behavior for x > 0.1 deviates from the observed
PO2 − x dependence. Furthermore, the aforementioned models can-
not predict the miscibility gab observed at temperatures below
685°C (see Fig. 12.3).
Lindemer71 modeled CeO2–x on the basis of a subregular solu-
tion model between CeO2 and Ce0.8O1.2. Two temperature-depend-
ent Redlich–Kister terms72 (i.e. regular and subregular solution
parameters) were found necessary in order to represent the non-
ideal behavior of the system. This approach also successfully pre-
dicted the low-temperature miscibility gap in CeO2–x. The success of
this model in reproducing the experimental data is shown in
Fig. 12.7. Hillert et al.73 applied a two-sublattice model to CeO2–x.
Regular and subregular solution parameters having a linear

b1469_Ch-12.indd 642 4/8/2013 12:40:55 PM


b1469 Catalysis by Ceria and Related Materials

Ceria and its Use in Solid Oxide Cells and Oxygen Membranes 643

Figure 12.7 Success of the subregular solution model employed by Lindemer71 in


reproducing experimental POeq - x isotherms for CeO2–x at various temperatures.
2
Reprinted with permission from Elsevier.

temperature dependence were again necessary to reproduce the


experimental data of Bevan and Kordis,38 including the miscibility
gap observed at low temperatures (below 685 °C). Yokokawa et al.74
also implemented a subregular solution model to describe the non-
stoichiometric CeO2–x phase in their thermodynamic calculations of
the ZrO2–YO1.5–CeO2–CeO1.5 system.

12.3.2.2 Influence of doping


The fluorite structure of CeO2–x forms extended solid solutions with
a large number of oxides of metals with valences ranging from 2+ to
6+. The influence of doping with metal cations of various valencies
on the thermodynamic properties of defect formation is discussed
here.
The substitution of Ce by isovalent cations, such as Zr4+, does
not necessitate the formation of additional (charge-compensating)

b1469_Ch-12.indd 643 4/8/2013 12:40:55 PM


b1469 Catalysis by Ceria and Related Materials

644 C. Chatzichristodoulou et al.

defects. Nevertheless, a number of investigations indicate enhanced


reducibility of CeO2–ZrO2 solid solutions.75–79 Chiodelli et al.80 found
a maximum in the electronic conductivity of CeO2–ZrO2 solid solu-
tions for the composition Ce0.7Zr0.3O2–x in line with the reported
enhanced reducibility of Ce. The enhanced reducibility of Ce on Zr
doping is related to the structural distortion/relaxation introduced
in the lattice due to the mismatch of Zr4+ in CeO2–x, as recent density
functional theory (DFT) calculations by Tang et al.81 indicate. DFT
calculations by Chen and Chang78 show a minimum for both the
∑∑
formation and migration energy of VO for the composition
Ce0.5Zr0.5O2–x. DFT calculations by Andersson et al.82 for (Ti, Hf, Zr,
∑∑
Th)-doped CeO2, also indicate facilitated VO formation upon
decreasing ionic radius of the isovalent dopant due to favorable
geometric relaxation of CeCe ¢ - VO∑∑ clusters, predicting a decreased
∑∑
tendency for VO formation on doping with the voluminous Th.
Steele and Floyd7 were the first to show that solid solutions of
ceria with rare earth oxides promote the reduction of Ce at tem-
peratures below 900 °C. Since then, a large number of investigations
on ceria doped with lower valent cations have verified this observa-
tion.42,43,75,83–89 The value of DH O2 ,x at x ≈ 0 has been found to
decrease (in absolute value) for increasing doping84,87,89 at least up
to 20 mol% of trivalent and 14 mol% of divalent doping.
As already pointed out previously, besides DH O2 ,x , the reducibil-
ity of ceria depends also on the configurational and nonconfigura-
tional entropy associated with its reduction. Therefore, DG O2 ,x
rather than DH O2 ,x should be compared for different compositions
to determine the effect of doping on the reducibility of ceria. Since
DG O2 ,x depends on temperature, any comparison should be per-
formed at different temperatures. Such a comparison is shown in
Fig. 12.8 for CeO2–x,38,39 Ce1–yCayO2–y–x,84 Ce1–2ySm2yO2–y–x,87 and Ce1–
2yGd2yO2–y–x,
86,90
at 500 °C, 1000 °C, and 1500 °C. As can be seen,
DG O2 ,x decreases (in absolute value) with increasing acceptor dop-
ing at 500 °C for x < 0.1, indicating that within this oxygen nonstoi-
chiometry range and at this temperature, acceptor-doped ceria is
more readily reduced than pure ceria. At 1000 °C, though, doping
has practically no effect on the reducibility of ceria. At 1500 °C the

b1469_Ch-12.indd 644 4/8/2013 12:40:55 PM


b1469 Catalysis by Ceria and Related Materials

Ceria and its Use in Solid Oxide Cells and Oxygen Membranes 645

Figure 12.8 Nonstoichiometry dependence of DG O2 ,x for CeO2–x,38,39 Ce1–yCay


O2–y–x,84 Ce1–2ySm2yO2–y–x,87 and Ce1–2yGd2yO2–y–x86,90 at 500 °C, 1000 °C, and 1500 °C.

situation is reversed, and most of the acceptor-doped ceria composi-


tions shown in Fig. 12.8 are less easily reducible than pure ceria.
Mogensen91 compared the composition dependence of DH O2 ,x
for pure and Gd-doped ceria, and concluded that the value of
DH O2 ,x is rather insensitive to the level of doping when the compari-
son is made at the same concentration of oxide ion vacancies.
A similar plot is shown here, Fig. 12.9, with the addition of data for
Sm- and Ca-doped ceria. The value of DH O2 ,x appears to be strongly
dependent on composition for both pure and doped ceria, but only
weakly dependent on doping type and level. This indicates that the
total concentration of oxide ion vacancies is the main factor affect-
ing the value of DH O2 ,x .
A particular case of dopants is that of Pr and Tb, which have the
potential to be in a mixed valence state 3+/4+ at more oxidizing
conditions than Ce as X-ray absorption near-edge structure (XANES)
and X-ray photoelectron spectroscopy (XPS) measurements have
revealed.17,92–96 The partial molar enthalpy of oxygen ( DH O2 ,x ) asso-
ciated with reduction of Pr or Tb appears to be strongly dependent
on doping level, degree of reduction, and co-doping. Knauth and

b1469_Ch-12.indd 645 4/8/2013 12:40:55 PM


b1469 Catalysis by Ceria and Related Materials

646 C. Chatzichristodoulou et al.

Figure 12.9 Composition dependence of the partial molar enthalpy of oxygen


for CeO2–x,39 Ce0.9Sm0.1O1.95–x and Ce0.8Sm0.2O1.9–x,87 Ce0.9Gd0.1O1.95–x and
Ce0.8Gd0.2O1.9–x,86,90 Ce0.9Ca0.01O1.99–x, Ce0.9Ca0.045O1.955–x, Ce0.9Ca0.07O1.93–x, Ce0.9Ca0.1O1.9–x,
and Ce0.8Ca0.14O1.86–x.84

Tuller97 estimated a value of −5.8 eV for Ce0.3Pr0.7O2–x. Fagg et al.98


determined a value of −1.5 eV for Ce0.8Pr0.2O2–x by coulometric titra-
tion at 800–950 °C. Chatzichristodoulou and Hendriksen99 observed
a linear dependence of DH O2 ,x on x for Ce0.8Pr0.2O2–x, varying from
−2.6 eV at x = 0 to −1.9 eV at x = 0.1, using coulometric titration and
thermogravimetric measurements at 600–900 °C. A similar observa-
tion was made by Bishop et al.100 for Ce0.9Pr0.1O2–x, where the value of
DH O2 ,x was found to vary between −3.8 eV and −3.35 eV in the
nonstoichiometry range x = 0–0.05.
Both DH O2 ,x and its linear dependence on x was seen to decrease
continuously (in absolute value) with increasing Tb content in the
ternary oxide Ce0.8Pr0.2–zTbzO2–x from −2.6 eV for z = 0 to −2.1 eV for
z = 0.2 at x = 0.17 The trivalent fraction of Pr or Tb in ceria increased
when co-doping with Zr,101,102 whereas that of Tb decreased when
co-doping with Ca.94 The tetravalent fraction of Tb increased with
increasing relative concentration of Pr, whereas the tetravalent frac-
tion of Pr decreased with increasing relative concentration of Tb.17
A trend toward stabilization of higher relative amounts of Tb4+ with

b1469_Ch-12.indd 646 4/8/2013 12:40:56 PM


b1469 Catalysis by Ceria and Related Materials

Ceria and its Use in Solid Oxide Cells and Oxygen Membranes 647

increasing Tb content in Ce1–zTbzO2–x was inferred from analysis of


X-ray diffraction (XRD) and Raman results.103 These observations
can be rationalized on the basis of minimization of the elastic lattice
strain energy associated with the size mismatch of the dopants to the
host lattice.17
The solubility of hydrogen in ceria and rare-earth-doped ceria
(Ce0.8M0.2O1.9, M = Y, La, Nd, Sm, Gd, Yb), according to reaction
(12.15), has been investigated by Sakai et al.104 using secondary ion
mass spectrometry (SIMS) after annealing in 0.03 atm of D2O and
pO2 = 0.02 atm at 880 °C and 990 °C. The detected proton solubility
in CeO2 was ∼0.003%, but increased by a factor of 10–100 on dop-
ing. The proton concentration increased with decreasing ionic radii
of the rare-earth dopant, ranging from 0.03% for Ce0.8La0.2O1.9 to
1% for Ce0.8Yb0.2O1.9. A similar study undertaken for (Zr1−x
Cex)0.8Y0.2O1.9 revealed a decreasing proton concentration with
increasing Zr content.105 Recently, Yokokawa et al.106 pointed out a
positive correlation between hole conductivity, proton solubility,
and thermodynamic activity of YO1.5 in Y-doped zirconia–ceria solid
solutions. Kim et al.107 reported that a significant amount of water
may be adsorbed on the surface of nanocrystalline ceria, independ-
ent of doping, even under a nominally dry atmosphere in the tem-
perature range 300–600 °C. They also concluded that incorporation
in grain boundaries and bulk appeared to be negligible.

12.3.2.3 Effect of crystallite size on defect formation: surface defects


The formation of oxide ion vacancies is expected to be facilitated at
surfaces and grain boundaries of ceria due to the lower coordination
number of oxide ions occupying such sites. The expected enhanced
reducibility of surfaces and interfaces is apparent in nanocrystalline
ceria due to the increased surface and grain-boundary density.
Indeed, a number of experimental results using various measuring
techniques provide unequivocal evidence for the enhanced
reducibility of ceria surfaces and interfaces. Early evidence came
from temperature-programmed reduction by H2 (H2-TPR) experi-
ments.108–110 An analysis of the H2-TPR results of Zimmer et al.111 by

b1469_Ch-12.indd 647 4/8/2013 12:40:56 PM


b1469 Catalysis by Ceria and Related Materials

648 C. Chatzichristodoulou et al.

Tschöpe112 concluded that the surface composition after reduction


at 600 °C in 5% H2/Ar was CeO0.8 (assuming a surface concentration
of oxide ions of 9.6 nm−2 as estimated by Bernal et al.110 for the (110)
surface of ceria). In situ XPS experiments of nanocrystalline
La-doped ceria have shown that approximately 50% of the surface
Ce ions are reduced to Ce3+ at 600 °C in H2.113 Enhanced reduction
in nanocrystalline ceria has also been verified both thermogravi-
metrically114 and by coulometric titration.115
Janvier et al.116 determined a value of −4.18 eV for the partial
molar enthalpy of oxidation ( DH O2 ,x ) of nanocrystalline ceria by
thermogravimetry and concluded that the oxide ion vacancies form
exclusively at the surface of the crystallites within the pO2 range
0.008–0.2 atm at 600 °C. Chiang et al.117 estimated a decrease (in
absolute value) in DH O2 ,x of nanocrystalline ceria (10 nm) by more
than 4.8 eV compared to that of microcrystalline CeO2–x, based on
bulk electrical conductivity data. A DH O2 ,x value of −3.68 eV
was determined by Hwang and Mason118 for nanocrystalline ceria
(15 nm) by thermopower measurements. Space-charge effects were
not considered in the determination of DH O2 ,x from the activation
energies of the electronic conductivity117 and the thermopower,118
which would result in a lower DH O2 ,x (absolute) value by ca. 1–2 eV.
Kim et al.119 analyzed electrical conductivity data for nanocrystalline
CeO2 (∼ 30 nm) with a Mott–Schottky space-charge model and
reported a DH O2 ,x value of ∼ −7.6 eV. Tschöpe112 calculated a decrease
(in absolute value) in DH O2 ,x by 4.6 eV for the surface as compared
to the bulk of ceria by fitting coulometric titration data for nanocrys-
talline ceria with a Gouy–Chapman space-charge model. In his analy-
sis, only fully ionized oxygen vacancies were taken into account while
the contribution of singly ionized or neutral oxygen vacancies as well
as adsorbed oxygen was neglected. Nevertheless, Porat et al.115 and
Kim et al.120 observed unusual pO2 exponents for the oxygen nonstoi-
chiometry of nanocrystalline cerium oxide by coulometric titration
measurements, and attributed their results to lower ionization states
of oxide ion vacancies or surface adsorption of oxygen in the first
case115 and lower ionization states of oxide ion vacancies and impu-
rity segregation in the second case.120

b1469_Ch-12.indd 648 4/8/2013 12:40:56 PM


b1469 Catalysis by Ceria and Related Materials

Ceria and its Use in Solid Oxide Cells and Oxygen Membranes 649

Suzuki et al.121 determined the crystallite size dependence of


DH O2 ,x by analyzing electrical conductivity measurements. They
obtained a value of ∼ −3.6 eV for CeO2−x for a crystallite size of 10
nm, neglecting the influence of space-charge effects. DH O2 ,x had
a value similar to that of microcrystalline ceria at a crystallite size of
∼ 100 nm.121 Kosacki et al.122 also reported an increasing concentra-
tion of point defects in the lattice of ceria with decreasing crystallite
size, following an analysis of Raman spectra on thin films of
nanocrystalline ceria. The defect concentration decreased fast with
increasing crystallite size and reached a value similar to that of
microcrystalline ceria at a crystallite size of ∼ 50–100 nm.122
Since the enhanced reducibility of nanocrystalline ceria is pri-
marily due to facilitated oxide ion vacancy formation at the surface,
one would expect the condition of the surface (segregation of impu-
rities and dopants at the surface, surface structure, etc.) to strongly
affect the reduction properties of nanocrystalline ceria. Indeed,
thermogravimetric measurements by Kim et al.120 on nanocrystalline
ceria prepared by two different synthesis routes demonstrate sub-
stantially different reduction behavior. They measured pO2 expo-
nents varying from 0 to −1/2 for differently prepared nanocrystalline
ceria powders, and explained these results on the basis of oxide ion
vacancies at different ionization states in addition to charged impu-
rities segregated onto the surface.120
Atomistic calculations support the observed reduction in forma-
tion energy of oxide ion vacancies at ceria surfaces. Sayle et al.123
calculated a decrease (in absolute value) in DH O2 ,x by 7.74 eV for
(111), 14.1 eV for (110), and 25.66 eV for the (310) surface of ceria
compared to the bulk value. Even smaller values were calculated by
Conesa.124 Reduced (in absolute value) DH O2 ,x values for ceria sur-
faces compared to the bulk have also been estimated by DFT calcula-
tions, which further verify the facilitated VO∑∑ formation on the (110)
compared to the (111) surface.125–128
As with microcrystalline ceria, the doping of nanocrystalline
CeO2–x with various cations (Zr, La, Y, Tb, Lu, etc.) has been shown
to improve reducibility and to enhance thermal stability (by hinder-
ing crystallite growth).129–136

b1469_Ch-12.indd 649 4/8/2013 12:40:56 PM


b1469 Catalysis by Ceria and Related Materials

650 C. Chatzichristodoulou et al.

Excess oxygen deficiency is also expected for grain boundaries,


again due to subcoordinated oxide ion sites, albeit at a different
level compared to surface sites. Indeed, an excess of oxide ion
vacancies and Gd in the grain-boundary core of polycrystalline
Ce0.8Gd0.2O1.9–x was revealed by electron energy loss spectroscopy
(EELS) measurements.137 The composition at the core was calcu-
lated to be Ce0.59±0.04Gd0.41±0.04O1.24±0.17 and the fraction of Ce3+ at the
core was estimated to be 70%,137 which points to the existence of a
significant net positive charge at the grain-boundary core.
The surface of ceria crystallites will also be affected by segrega-
tion of impurities and dopants driven by (i) the electrostatic interac-
tion between charged solute ions (dopants or impurities) and the
space-charge potential, (ii) the elastic strain energy due to the size
and charge misfit of the solute ion in the ceria lattice, and (iii)
dipole interactions between solute-vacancy dipoles and the electric
field in the space-charge layers.138 The segregation of solute ions,
which also carry a net charge, will contribute to the overall charge
at the interface and the formation of the space-charge potential.
Experimental evidence of solute segregation in ceria is available.
CezPr1–zO2–x mixed oxides (30–40 nm), studied by Raman spectros-
copy and XPS, were found to be strongly enriched in Pr at the sur-
face. The thickness of the Pr-rich surface layer was found to increase
with increasing Pr content.139 Segregation of Pr at the surfaces and
grain boundaries of nanocrystalline Pr-doped ceria has also been
observed with SIMS and XPS analysis by Borchert et al.,95 and the
degree of segregation was found to increase after reduction. Strong
segregation of Gd at the grain-boundary core of Gd-doped ceria was
revealed by Lei et al. 137 using EELS (see previous paragraph) and by
Blom and Chiang140 by scanning transmission electron microscopy-
energy dispersive spectroscopy (STEM-EDS). Since the size mis-
match of Gd in CeO2–x is small (leading to a segregation energy of
only 0.05 eV112), the observed segregation of Gd at the grain bound-
ary is expected to be mainly due to the grain-boundary excess of
oxide ion vacancies and the subsequent positive space charge accu-
mulated at the grain-boundary core.

b1469_Ch-12.indd 650 4/8/2013 12:40:56 PM


b1469 Catalysis by Ceria and Related Materials

Ceria and its Use in Solid Oxide Cells and Oxygen Membranes 651

Another interesting consequence of the increased surface and


grain-boundary density of nanocrystalline ceria is the enhanced
solubility of dopants with a large size mismatch as they can be
accommodated at the interfaces. The apparent solubility of copper
in nanocrystalline CeO2–x was determined to be ∼ 10 mol%,141 which
is a significant enhancement relative to microcrystalline ceria, where
a solubility limit of ~ 1 mol% is expected (see Section 12.2.2.2).
Due to the large number of surface defects, significant defect
interaction is expected to take place, potentially resulting in defect
clustering and ordering. This topic has recently been investigated
for the (111) surface of ceria (which is the low energy surface of
ceria123,124) by several groups, using scanning tunneling microscopy
∑∑
(STM). Nörenberg and Briggs142 observed VO clusters, the domi-
∑∑
nant type being of triangular shape (containing three surface VO )
at room temperature. They also observed line defect clusters at
500 °C. Namai et al.143 found both line defects and triangular defects,
the line defects being dominant after annealing at 900 °C. This is
further supported by more recent results from Esch et al.144 Local
reconstruction takes place, with the oxide ions surrounding the
∑∑
multiple defects being displaced toward the VO sites, which is not
143
observed for point defects.
O2-electron paramagnetic resonance (EPR) and adsorbed
methoxy infrared (IR) experiments have shown that Tb promotes
∑∑
the generation of clustered VO defects at the surface of Tb-doped
∑∑
ceria nanoparticles. Crozier et al.145 observed VO ordering after
103

reduction of ceria nanocrystals above 730 °C by environmental


transmission electron microscopy (ETEM) and the formation of a
superlattice. The vacancy ordering was completely reversible by
reduction and re-oxidation with changing temperature and oxygen
partial pressure. They also observed an irreversible reconstruction
of the surface after reduction at 700 °C. The original sawtooth sur-
face made up of (111) nanofacets had smoothed into (110) ter-
races. They argued that this reconstruction may occur to avoid the
creation of a net dipole moment perpendicular to the surface after
reduction.145

b1469_Ch-12.indd 651 4/8/2013 12:40:56 PM


b1469 Catalysis by Ceria and Related Materials

652 C. Chatzichristodoulou et al.

12.4 Ionic and Electronic Conductivity of Pure


and Doped Ceria
12.4.1 Pure ceria
12.4.1.1 Ionic conductivity
CeO2−x is both an n-type semiconductor and an oxide ion conductor,
i.e. a mixed ionic and electronic conductor (MIEC).4,62,146–149 The
ionic transport takes place via hopping of oxide ions to vacant sites,
and the ionic conductivity σi may be expressed as:

s i = ÈÎ VO∑∑ ˘˚ e mi
(12.28)

where e is the electron charge and µi the ion mobility. The ionic
conductivity of pure reduced ceria at 1000 °C is less than 3% of
the total conductivity at a pO2 of 10−6 atm and even lower at lower
pO2 and at lower temperatures.16 The ionic conductivity shows a
maximum with increasing oxide ion vacancy concentration,
which at 1000 °C is estimated to be ca. 7 × 10−2 S⋅cm−1 for CeO1.9.16
The total conductivity (ionic and electronic) for this condition is
2.5 S⋅cm−1.16 The oxide ion mobility µi follows an Arrhenius
equation:

mi0 Ê H ˆ
mi = exp Á - i ˜
T Ë kT ¯
(12.29)

where k is the Boltzmann constant, Hi is the activation energy


for oxide ion hopping and µi0 the pre-exponential factor. Tschöpe
et al.150 estimated the values of 0.71 eV and 1.5 × 102 cm2⋅kV−1⋅s−1 for
Hi and µi0, respectively at pO2 = 10−4 atm and 350–650 °C (for CeO2
with 2000 ppm maximum total metallic impurities). The capture of
oxygen vacancies by association (discussed later) is implicitly taken
into account in the estimation of Hi and µi0, which therefore account
for the effective oxide ion mobility.

b1469_Ch-12.indd 652 4/8/2013 12:40:57 PM


b1469 Catalysis by Ceria and Related Materials

Ceria and its Use in Solid Oxide Cells and Oxygen Membranes 653

12.4.1.2 Electronic conductivity


n-type conductivity takes place by small polaron transport.45,63,151
A small polaron is a defect created when an electronic carrier
becomes trapped at a given site as a consequence of the displacement
of adjacent atoms or ions in the crystal lattice. The entire defect, i.e.
the electron (or electron hole) plus distortion, migrates by thermally
activated hopping. The small polaron in CeO2−x may be identified as
Ce′Ce . The electronic conductivity σe may be expressed as:

s e = ÎÈCeCe
¢ ˚˘ e me
(12.30)

where µe is the small polaron mobility, which has the form:

me0 Ê H ˆ
me = exp Á - e ˜
T Ë kT ¯
(12.31)

where He is the activation energy for the small polaron hopping. The
pre-exponential factor µe0 is given by:

ÈCexCe ˘
me0 = Î ˚ ea 2n 0 k
ÈCeCe ˘ + ÈCeCe ˘
˚ Î ¢ ˚
x
Î (12.32)
where α is the lattice parameter, ν0 the phonon frequency, and the
Ce′Ce fraction represents the fraction of sites that are available for
hopping of the small polaron.
The variation of conductivity σ of CeO2−x at a fixed temperature
(1000 °C) is shown as a function of x in Fig. 12.10. Even though σe
is known to be proportional to x in the vicinity of stoichiometry (up
to x ≈ 10−2) saturation occurs at x ≈ 0.04 and σe has a broad maxi-
mum with a peak value at x ≈ 0.1. The activation energy of the small
polaron mobility He varies with x from about 0.2 eV near stoichiom-
etry to about 0.5 eV at a vacancy concentration of x = 0.2.45,63,152 The
increase in He with increasing x is attributed to defect interactions
at higher x values, as well as to the increase in lattice parameter
with increasing x.45,63,153,154 Tuller and Nowick45 further concluded

b1469_Ch-12.indd 653 4/8/2013 12:40:57 PM


b1469 Catalysis by Ceria and Related Materials

654 C. Chatzichristodoulou et al.

Figure 12.10 Conductivity of CeOy at 1000 °C as a function of nonstoichiometry.


Data are from Tuller and Nowick45 (O) with earlier data from Blumenthal et al.6 (x).
After Tuller and Nowick.45 Reprinted with permission from Elsevier.

that the fraction of the effective number of electron carriers


decreases with increasing x for x > 0.04, based on Seebeck coeffi-
cient measurements. The increase in He, along with the decrease in
the fraction of effective carriers as x increases, accounts for the fact
that σ(x) is not a symmetrical function about x = 0.25, as predicted
from the pre-exponential factor of conductivity, but has a broad
maximum shifted to lower x as shown in Fig. 12.10. A comparison
of polycrystalline and single-crystal conductivity data indicate
that the grain boundaries do not form any barrier to electron
migration.155
Electron hole (p-type) conductivity has been observed in ceria at
pO2 ≈ 1–400 atm in the temperature range 600–1100 °C.61,156 Electron
holes were envisaged to arise according to reaction (12.16), where
the vacancies are induced by lower valent impurities61,156 or by for-

mation of cation vacancies VCe having an effective negative charge
ν. A pO2 exponent of 1/4 was observed at 700 °C and pO2 ≈ 400
156

atm.156 An electron hole mobility of approximately 10−5 cm2⋅V−1⋅s−1


was estimated at 800 °C, significantly lower than the electron mobil-
ity of about 6.1 × 10−3 cm2⋅V−1⋅s−1 at the same temperature.61

b1469_Ch-12.indd 654 4/8/2013 12:40:57 PM


b1469 Catalysis by Ceria and Related Materials

Ceria and its Use in Solid Oxide Cells and Oxygen Membranes 655

12.4.1.3 Influence of crystallite size


Chiang et al.117 investigated the crystallite size dependence of the
electrical conductivity of CeO2 over a broad temperature and pO2
range. At 600 °C and pO2 ≈ 1 atm the bulk conductivity of CeO2
nanocrystals (with a crystallite size of ~ 10 nm) was ~ 104 times
higher than the extrapolated electronic conductivity of thermally
coarsened counterparts (~ 5 µm grain size).117 They attributed their
findings to the lowered DH O2 ,x (absolute) value observed for the
nanocrystalline material, associated with facilitated VO•• and Ce′Ce for-
mation at grain-boundary sites and the consequent increase in the
concentration of ionic and electronic charge carriers.117
Tschöpe et al.150 confirmed that nanocrystalline specimens
(~ 26 nm) exhibited electronic conductivity under conditions at
which microcrystalline samples showed impurity-controlled ionic
conductivity. The electronic conductivity of nanocrystalline CeO2
was enhanced relative to the intrinsic electronic conductivity of pure
single crystalline CeO2 and increased with decreasing grain size. The
opposite trend was observed for the partial ionic conductivity. The
loss of impurity-controlled ionic conductivity was also observed in
0.2 at% Gd-doped samples, which is indicative of a significant space-
charge effect related to a positive grain-boundary excess charge.
This situation results in the depletion of oxide ion vacancies and the
accumulation of electrons and acceptor impurities in the space-
charge layer, and consequent depletion of oxide ion vacancies from
the bulk. Indeed, measurements of the electronic structure of
cerium oxide (111) surfaces by ultraviolet photoelectron spectros-
copy (UPS) revealed the existence of a surface dipole with positive
excess charge in the interface and a negative space-charge layer.157
An analytical model was further derived by Tschöpe158 (under
the assumptions of (i) large grain sizes compared to the screening
length, (ii) a brick-layer model topology, (iii) thermodynamic
equilibrium of all point defects (the Gouy–Chapman situation),
and (iv) defect thermodynamic parameters for single crystalline
CeO2 applied to the bulk of the nanoparticles) to investigate the
effect of the space-charge potential on the partial conductivities of

b1469_Ch-12.indd 655 4/8/2013 12:40:57 PM


b1469 Catalysis by Ceria and Related Materials

656 C. Chatzichristodoulou et al.

polycrystalline CeO2. The predictions of this space-charge model


(i.e. (a) a transition from ionic- to electronic-dominated conduc-
tivity with decreasing grain size, (b) an enhanced electronic con-
ductivity of nanocrystalline compared to single crystalline CeO2,
and (c) a decreasing/increasing activation energy of the partial
electronic/ionic conductivity with decreasing crystallite size due to
the temperature dependence of the space-charge effect) are in
qualitative agreement with experimental results.117,118,150
Kim and Maier159 investigated the electrical conductivity of 0.15
at% Gd-doped and nominally pure nanocrystalline CeO2 (~ 30 nm)
by impedance spectroscopy and Hebb–Wagner polarization with an
electronically blocking electrode. They showed that a Mott–Schottky-
type space-charge model can quantitatively account for all their
experimental results. The space-charge potential was found to be
∆φ(0) ≈ 0.3 V, being only weakly dependent on temperature
∂Df(0) ∂Df(0)
(T1 ∂1/T ª 0.1 V) and oxygen partial pressure ( ∂P ª 0). This leads to
O2
••
VO depletion at the space-charge regions hindering the grain-to-
grain ionic transport, and electron accumulation leading to highly
electronically conducting grain-boundary layers that bypass elec-
tronic transport through the bulk. The activation energy for ionic
conduction in the bulk of 0.15 at% Gd-doped ceria was Hi,∞ = 0.73 eV,
close to that of microcrystalline counterparts, whereas a much
higher value of Hi,⊥ = 1.5 eV was determined for the activation energy
for ionic transport through grain boundaries, as expected for a posi-
tive space-charge potential according to the relation:159

Ê 1 ∂Df (0) ˆ
H i ,^ = H i ,• + 2e Á Df (0 ) + (12.33)
Ë T ∂1 T ˜¯

The electron accumulation at space-charge regions was found to


dominate the conducting properties of nominally pure CeO2, with a
temperature dependence (in the extrinsic regime) given by:159

k
(
∂ ln s e ,||T ) = DH O2 ,x Ê
- H e + e Á Df (0) +
1 ∂Df (0) ˆ
(12.34)
∂1 T 4 Ë T ∂1 T ˜¯

b1469_Ch-12.indd 656 4/8/2013 12:40:57 PM


b1469 Catalysis by Ceria and Related Materials

Ceria and its Use in Solid Oxide Cells and Oxygen Membranes 657

A misprint in the corresponding Eq. (21) in Kim et al.119 has been


corrected here (ln(σe,||T ) rather than ln (σe,||)), and the activation
energy of the electronic conductivity in the bulk Ee,∞ has been sub-
stituted by the partial molar enthalpy of oxidation and the migra-
tion enthalpy, using the relation: Ee ,• = DH O2 ,x - H e . In Eq. (12.34),
4
the value of DH O2 ,x corresponds to the bulk part of the nanocrys-
tals. They observed a value of −1.9 eV, which for He ≈ 0.4 eV,
results in DH O2 ,x ª -7.6eV , somewhat lower (in absolute value)
than that of microcrystalline CeO2. They also concluded that in
nominally pure CeO2 the space-charge distribution is inhomoge-
neous, leaving large contact areas almost unaffected for ionic
transport.
Rupp and Gauckler160 and Mansilla et al.161 confirmed the
increasing total DC conductivity with decreasing average crystallite
size for spray-pyrolyzed and ion-beam-assisted electron beam evapo-
ration deposited CeO2 thin films respectively. Hwang and Mason118
estimated an activation energy of 1.6 eV for ionic transport in
nanocrystalline ceria (15 nm), more than twice the value of micro-
crystalline ceria. Litzelman et al.162 also reported an activation energy
of 1.58 eV for ionic transport in nanocrystalline (25–45 nm) ceria
thin films, determined by Hebb–Wagner electron blocking measure-
ments. On the basis of the Mott–Schottky space-charge model pro-
posed by Kim and Maier,159 these values correspond to a space-charge
potential of ca. 0.35 eV (assuming Hi,∞ ≈ 0.7 eV). Litzelman et al.162
also determined the activation energy of the electronic conductivity
to be −1.33 eV, which for a space-charge potential of 0.35 eV, would
result in DH O2 ,x ª -5.52 eV , significantly lower (in absolute value)
than that of microcrystalline ceria (see Fig. 12.5 in Section 12.3.2.1).
This value is also significantly lower than that obtained by Kim and
Maier159 and it therefore remains unclear as to what extent the ther-
modynamics of the bulk portion of nanocrystalline CeO2−x differ
from the microcrystalline state.
Litzelman et al.163 attempted to modify the space-charge potential
of ceria nanocrystals by controlled in-diffusion of cations. They
observed a decrease in electrical conductivity by more than one order
of magnitude, while the modified films remained predominantly

b1469_Ch-12.indd 657 4/8/2013 12:40:57 PM


b1469 Catalysis by Ceria and Related Materials

658 C. Chatzichristodoulou et al.

electronically conducting. This was attributed to a decrease of the


space-charge potential to a lower yet still positive value, resulting in a
decrease in electron concentration and an increase in oxide ion
vacancy concentration in the space-charge layers.
Recently, Kim et al.164 observed abnormally fast oxygen ionic
conduction (about two orders of magnitude higher than that of
Ce0.8Gd0.2O1.9) in polycrystalline 1D CeO2 nanowires. Their investiga-
tions indicated that fast ionic conduction is related to enhanced ionic
mobility, rather than increased charge carrier concentration. They
speculated that an amorphous layer that encapsulates the nanowire
may serve as a fast conduction pathway for ionic transport.

12.4.2 Effect of doping on the ionic and electronic conductivity of ceria


12.4.2.1 Ionic conductivity of doped ceria
Whereas the ionic conductivity is always much lower than the elec-
tronic conductivity in pure reduced ceria, the situation is quite dif-
ferent in ceria doped with oxides of two- or three-valent metals due
to the introduction of oxide ion vacancies as charge compensating
defects, according to Eqs. (12.2) and (12.3). The electronic conduc-
tivity in air may be very low and under these conditions doped ceria
is an excellent electrolyte.
The total ionic conductivity consists of a contribution from the
bulk of the grains and one from the grain boundaries, the two con-
tributions being in series. The specific grain-boundary conductivity
of acceptor-doped ceria is at least two orders of magnitude lower
than the corresponding bulk value.165–189 This is mainly due to the
formation of impurity phases at the grain boundary, which block
ionic transport168–170 and due to the depletion of oxide ion vacancies
in the space-charge layer. In materials of normal or low purity, the
effect of the grain-boundary impurity phase is dominant, whereas in
materials of high purity, the effect of V•• O depletion at the space-
charge layer is dominant.
SiO2 is one of the major impurities present in CeO2 ceramics,
along with alkali and transition metal oxides. During sintering,

b1469_Ch-12.indd 658 4/8/2013 12:40:57 PM


b1469 Catalysis by Ceria and Related Materials

Ceria and its Use in Solid Oxide Cells and Oxygen Membranes 659

these impurities accumulate or disperse at grain boundaries, and


react with each other and the bulk components to form an inter-
granular siliceous phase. The composition, location, viscosity, and
wetting properties of the siliceous phase vary with the type and
amount of impurities in the starting powders, sintering atmosphere
and temperature, and cooling rate. The grain-boundary conductiv-
ity is thus observed to vary significantly under different conditions.
Guo and Waser190 concluded that the ionic transport across the
grain boundaries occurs solely through grain-to-grain contacts; the
grain-boundary impurity phase thus contributes to the grain-bound-
ary blocking effect by decreasing the conduction path width and
constricting current lines.
In materials of high purity in which the siliceous phase is not
observed, the specific grain-boundary conductivity is still at least two
orders of magnitude lower than that of the bulk.171,191 This is attrib-
uted to the depletion of V•• O in the grain-boundary space-charge
layer.150,158,159,191 In this case, the activation energy for the grain-
boundary conductivity is higher than that for the bulk.166,167,175,178
The grain-boundary blocking effect disappears at high enough
temperatures183 (> 600 °C) or high enough dopant concentration166–168
(> 15 mol% CaO). At high temperatures, the effective segregation
of acceptor dopants and consequent masking of the core charge,
leads to minimized oxide ion vacancy depletion in the space-charge
layer. Furthermore, the width of the space-charge layer decreases
with increasing dopant content (being as small as 0.1 nm for high-
dopant concentrations). The very small space-charge layer width
and the modest depletion of oxide ion vacancies in the space-charge
layer account for the negligibly small grain-boundary resistance at
high temperatures and high dopant concentrations.
A large number of studies of the ionic conductivity of ceria
doped with alkaline and rare-earth metal oxides have been repor
ted.16,28,83,172,173,175,181,182,185,186,189,192–213 On the basis of Eq. (12.28) one
would expect a linear relation between σi and acceptor doping (for
moderate acceptor doping, i.e. when [OxO ] ª 2 ). In reality this holds
192,201
true only for V•• O concentrations below ca. 1–3%. For higher
acceptor dopant contents the lattice ionic conductivity σi,∞ exhibits a

b1469_Ch-12.indd 659 4/8/2013 12:40:57 PM


b1469 Catalysis by Ceria and Related Materials

660 C. Chatzichristodoulou et al.

Figure 12.11 Lattice ionic conductivity at 400 °C as a function of acceptor dopant


concentration for different rare-earth elements. Calculated from activation energy
and pre-exponential factor data.201

broad maximum as shown in Fig. 12.11. The maximum is consist-


ently observed for a nominal V•• O concentration of
172,175,182,185,192,194,197–199,201
~ 2–5% and is accompanied by a minimum of
the activation energy (at a somewhat different V•• O concentration, as
the pre-exponential factor is also a function of dopant content) as
shown in Fig. 12.12.
These phenomena are attributed to competitive interactions
between point defects and the formation of defect associates.194,214,215
The decrease in activation enthalpy in the dilute range is due to a
decrease in association enthalpy as a result of electrostatic interac-
tions between the associated pair and the unassociated acceptor cati-
ons. The ensuing decrease in conductivity and increase in activation
energy with increasing dopant concentration is ascribed to the
development of deep vacancy traps as a result of defect clustering
and the formation of microdomains. For associated vacancies the
activation energy Hi consists of a migration Hm and an association
(dissociation) Ha term: Hi = Hm + Ha .
Maximum lattice ionic conductivities in the range of 0.04–0.06
S⋅cm−1 at 800 °C have been reported with alkaline-earth doping, with

b1469_Ch-12.indd 660 4/8/2013 12:40:57 PM


b1469 Catalysis by Ceria and Related Materials

Ceria and its Use in Solid Oxide Cells and Oxygen Membranes 661

Figure 12.12 Activation energy as a function of acceptor dopant concentration for


different rare-earth elements. Inset shows the minimum activation enthalpy as a
function of dopant ionic radius. Adapted from Faber et al.201 Reprinted with kind
permission from Springer Science and Business Media.

Ca and Sr being the optimum dopants.192,196–198 Higher lattice ionic


conductivities have been achieved when doping with rare-earth ele-
ments, reaching values of ca. 0.1 S⋅cm−1 at 800 °C for Gd or Sm dop-
ing168,183,200–202,204,213,216 The maximum ionic conductivities vary within
a factor of ca. 10 at 400 °C201,202,208 and ca. 3 at 800 °C for different
rare-earth dopants.200
The position and value of the minimum of the activation energy
Hi as a function of rare-earth dopant content appears to vary with
dopant type. In fact the minimum value of the activation energy
Hi,min shows a smooth parabolic-type dependence with the ionic radii
of the dopant, exhibiting a minimum in the region between Gd
(1.053 Å) and Nd (1.109 Å),168,175,195,200–202,208 as shown in the inset of
Fig. 12.12. The minimum in activation energy and maximum in lat-
tice ionic conductivity, observed for Gd (1.053 Å) or Sm (1.079 Å)
doping, was attributed to the minimization of the disturbance to the
lattice of ceria achieved for these dopants.217 Kim21 introduced the
concept of the matching ionic radii rm i.e. the ionic radii of a dopant

b1469_Ch-12.indd 661 4/8/2013 12:40:58 PM


b1469 Catalysis by Ceria and Related Materials

662 C. Chatzichristodoulou et al.

that causes neither expansion nor contraction of the host lattice,


and suggested an rm value of 1.038 Å for ceria, close to that of Gd
(see Section 12.2.2.2).
Enhanced ionic conductivity (compared to that of singly doped
counterparts in the respective investigations) has been achieved in
a number of co-doped ceria systems, namely in Sm/Gd,218 Nd/Lu,219
Ca/Y,220 Gd/Y,221,222 Nd/Sm207,210,223 Sr/Y,224 and Nd/Gd213 co-doped
ceria. This is attributed to (i) increased configurational entropy in
the cation sublattice, leading to suppressed association of oxide ion
vacancies225 and (ii) reduced global lattice strain (quantified by the
deviation of the lattice parameter from that of CeO2). Examples of
decreased ionic conductivity after co-doping have been reported,
though, for La/Yb,173 Sm/Y,173 and La-Nd/Gd-Y226 combinations.
These results demonstrate that minimization of the global lattice
strain is not an adequate criterion in terms of improving lattice ionic
transport, and highlight the importance of the local lattice strain.
Extensive atomistic simulation work227–232 has provided valuable
insights into the nature of defect association and migration. It is
generally recognized that both attractive coulombic interactions and
attractive or repulsive elastic lattice relaxation play a major role in
the energetics of ionic transport. The atomistic simulations con-
clude that for small dopants, the oxide ion vacancies preferentially
associate with dopant ions, whereas for large dopants the oxide ion
vacancies tend to associate with Ce4+ host cations.228-–231 The switcho-
ver is estimated to occur at a rare-earth dopant ionic radius similar
to that of Gd,228 which offers a rational for the minimum in activa-
tion energy and the maximum in ionic conductivity observed for
Gd- or Sm-doped ceria.168,183,200–202,204,213,216 Andersson et al.229 con-
cluded that the minimum association energy is reached when the
best balance between repulsive elastic and attractive coulombic
interactions is obtained, which according to their DFT calculations
happens for Pm or Sm. They estimated that the migration enthalpy
in the absence of rare-earth dopants around the oxide ion vacancy
is 0.46 eV, in agreement with nuclear magnetic resonance (NMR)
results, which have yielded a value of 0.49 eV.233 They also concluded
that the migration enthalpy depends strongly on the initial cationic

b1469_Ch-12.indd 662 4/8/2013 12:40:58 PM


b1469 Catalysis by Ceria and Related Materials

Ceria and its Use in Solid Oxide Cells and Oxygen Membranes 663

configuration around the oxide ion vacancy and the migration path,
being larger for migration paths in between dopant cations.229
A similar conclusion was recently reached by Frayret et al.231 This is
in agreement with experimental results showing an increasing acti-
vation energy at higher acceptor dopant contents (see Fig. 12.12).
Recent micro-calorimetric measurements by Navrotsky et al.234
on ceria doped with Y, Gd, or La, have shown that the oxide ion
vacancies tend to occupy the first nearest neighbor sites to Y and Gd
dopants, but also to La, in contrast to estimates from atomistic simu-
lations.228–231 A coincidence between the dopant content where the
maximum in ionic conductivity occurs and that in enthalpy of for-
mation (from the component binary oxides) for both La- and
Gd-doped ceria was also observed.235 A rather sharp transition from
a situation of negligible clustering at a low doping level to almost
complete clustering at high (above the conductivity maximum) was
detected. Higher lattice ionic conductivity was recently observed for
Nd-doped ceria compared to Gd- or Sm-doped ceria,208 further com-
plicating the long-lasting dispute as to whether Gd or Sm is the
optimum dopant in terms of lattice ionic conductivity. Furthermore,
a clear dependence of the transport properties of rare-earth-doped
ceria (containing a low level of impurities) on sintering temperature
and preparation method is well established.178,209,236,237 The lattice
conductivity of Y-doped ceria was found to vary by more than one
order of magnitude depending on sintering temperature.178
These results indicate that the full picture of ionic transport in
doped ceria is still not complete and that there are additional
parameters in play, such as deviations from the assumption of ran-
domly distributed dopants, which may be influenced by powder
synthesis and processing. High-resolution transmission electron
microscopy (HRTEM), selected area electron diffraction (SAED),
and EELS investigations have revealed the existence of dopant-rich
nano-sized domains in ceria doped with La,238 Sm,239 Y,240 Yb,241
Ho,242 Dy,243 Tb,244 and Gd.245 Ordering of oxygen vacancies was
observed to take place in these domains. Clustering of dopants and
oxide ion vacancies was also indicated by analysis of extended X-ray
absorption fine structure (EXAFS) measurements for Y- and

b1469_Ch-12.indd 663 4/8/2013 12:40:58 PM


b1469 Catalysis by Ceria and Related Materials

664 C. Chatzichristodoulou et al.

La-doped ceria.246 A correlation between synthesis process, nanodo-


main size, and total ionic conductivity was observed for Sm,239 Yb,241
Ho,242 and Dy243-doped ceria. Furthermore, recent results question
the established solubility limits of rare-earth cations in ceria.
Although the lattice parameter of ceria increases with La doping
up to 40 mol% LaO1.5,26,247 the true solubility turns out to be below
20 mol% LaO1.5 as evidenced from the formation of C–La2O3 after
annealing Ce0.8La0.2O1.9 at 1000 °C for 7 days.18 Pronounced devia-
tions from Vegard’s law have been consistently observed for the
lattice constant of Gd-doped ceria.26,185,248 These may be rationalized
by the formation of C–Gd2O3 domains within the ceria fluorite
lattice.185
Little has been reported with respect to the aging of doped ceria
electrolytes. Significant degradation of both the bulk and grain-
boundary ionic conductivity was observed in ceria doped with more
than 20 mol% YO1.5 (reaching 10%, and 15% for Ce0.7Y0.3O1.85) after
annealing at 1000 °C for 8 days.186 Even more pronounced degrada-
tion was detected for ceria doped with more than 15 mol% GdO1.5
(reaching ca. 50% and 60% for Ce0.7Gd0.3O1.85) under the same
annealing conditions.249 These observations are attributed to the
formation of C–Y2O3 or C–Gd2O3 microdomains in the ceria
lattice.186,249

12.4.2.2 Electronic conductivity of doped ceria


As for pure CeO2−x, electronic conductivity is introduced in doped
ceria by increasing temperature and decreasing pO2, due to the
reduction of Ce4+ to Ce3+ and the formation of electronic defects.
The n-type conductivity takes place via a small polaron mecha-
nism.83,250 According to Eqs (12.30), (12.31), and (12.32), the elec-
tronic conductivity of doped ceria is a function of the degree of
reduction x, the electronic migration enthalpy He, and the pre-
exponential term of the electron mobility µe0.
As discussed in Section 12.3.2, dealing with the defect thermody-
namic properties of doped ceria, acceptor doping enhances the
reducibility of ceria at temperatures below 1000 °C. Therefore doped

b1469_Ch-12.indd 664 4/8/2013 12:40:58 PM


b1469 Catalysis by Ceria and Related Materials

Ceria and its Use in Solid Oxide Cells and Oxygen Membranes 665

ceria has a somewhat higher concentration of electronic defects than


pure ceria at the same temperature (below 1000 °C) and pO2. This
effect becomes increasingly more pronounced on increasing the
dopant content up to 20 mol% for trivalent and 14 mol% for diva-
lent doping, but is practically independent of dopant type.
As with pure CeO2−x, the migration enthalpy for electronic con-
duction increases with increasing concentration of oxide ion vacan-
cies (i.e. acceptor dopant content).250,251 Levy et al.250 determined a
migration enthalpy of 0.33 eV for Ce0.99Y0.01O1.995−x and 0.55 eV for
Ce0.74Y0.26O1.87−x at x ≈ 0.01. On the other hand, the migration
enthalpy for electronic conduction does not depend on dopant
type.252 Somewhat lower migration enthalpy values were determined
by Xiong et al.251 for Ce1−yYyO2−y/2−x. The migration enthalpy increased
from ca. 0.2 eV for y = 0.1 to ca. 0.35 eV for y ≥ 0.2.251 It should be
pointed out that Xiong et al.251 derived apparent migration enthalpy
values using an Arrhenius expression without an inverse tempera-
ture pre-exponential factor. Correcting for this would raise the
migration enthalpy values by ca. 0.1 eV, resulting in good agreement
with the values estimated by Levy and Fouletier.250
According to Eq. (12.31), the pre-exponential factor is expected
to decrease with increasing dopant content due to the decreasing
number of sites where the small polaron may jump. This was dem-
onstrated to be the case in Ce1−yGdyO2−y/2−x and Ce1−yCayO2−y−x by
Mogensen et al.16 As with the concentration of electronic defects and
the electronic migration enthalpy, the pre-exponential factor of the
electronic mobility is also found to be rather insensitive to dopant
type in 20 mol% REO1.5 doped ceria.252
In contrast to the ionic conductivity, the magnitude of the elec-
tronic conductivity is rather insensitive to the ionic radii of the
dopant.16,252,253 A slightly increasing tendency has been observed
with increasing dopant ionic radii from Y to La.252 The magnitude
of the electronic conductivity is generally observed to decrease with
increasing doping.16,251 Mogensen et al.16 observed a decreasing
electronic conductivity with increasing dopant content in the sys-
tems Ce1−yGdyO2−y/2−x and Ce1−yCayO2−y−x at 1000 °C. Xiong et al.251
observed a decreasing electronic conductivity with increasing Y

b1469_Ch-12.indd 665 4/8/2013 12:40:58 PM


b1469 Catalysis by Ceria and Related Materials

666 C. Chatzichristodoulou et al.

content in the system Ce1−yYyO2−y/2−x for 0.1 ≤ y ≤ 0.4, despite the


increasing concentration of electronic defects with increasing Y
content (arising from the decreasing (absolute) value of DG O2 ,x
with increasing y (see Fig. 12.8 in Section 12.3.2.2)). This is partly
due to the increasing migration energy with increasing y from 0.1
to 0.2, but mainly due to the decreasing value of the pre-exponen-
tial factor upon increasing dopant content. A similar trend was
witnessed in Ce1−ySmyO2−y/2−x at 700–950 °C.254 According to Steele216
and Wang et al.,204 the electronic mobility of Ce0.8Gd0.2O1.9−x is lower
than that of Ce0.9Gd0.1O1.95−x. Nevertheless, the conductivity of
Ce0.8Gd0.2O1.9−x is higher than that of Ce0.9Gd0.1O1.95−x at tempera-
tures between 500 °C and 800 °C due to the higher concentration of
electronic defects in Ce0.8Gd0.2O1.9−x216 (arising from the decreasing
(absolute) value of DG O2 ,x with increasing Gd content (Fig. 12.8)).
The situation is reversed at higher temperatures due to the fact
that the Gd content dependence of DG O2 ,x is reversed (Fig. 12.8).
Navarro et al.255 estimated similar electronic conductivities for
Ce0.9Gd0.1O1.95−x and Ce0.8Gd0.2O1.9−x in the temperature range
800–1000 °C.
The electrolytic domain (ti ≥ 0.99) of doped ceria is limited at
low pO2 due to the onset of electronic conductivity on reduction of
Ce4+. The low pO2 limit of the electrolytic domain depends on
dopant type and doping level, and shifts to higher pO2 with increas-
ing temperature. The width of the electrolytic domain is maximized
for 10–20 mol% Gd- or Sm-doped ceria,200 due to the optimized total
ionic conductivity achieved in these compounds. The low pO2 limit
of these compounds shifts from ca. 10−29 atm at 400 °C and ca. 10−16
atm at 600 °C to ca. 10−9 atm at 800 °C.193,216,256,257 For comparison, the
low pO2 limit for Ce0.9Y0.1O1.95−x was determined to extend down to a
pO2 of about 10−13 atm at 600 °C.83
Significantly enhanced electronic conductivity at elevated oxy-
gen partial pressures has been observed in ceria doped with Pr or
Tb compared to other rare-earth or alkaline-earth elements.98,100,258–271
This is attributed to the contribution of small polarons in the Pr or
Tb 4f band. The electronic conductivity of Ce0.8Pr0.2O2−x is higher
than that of Ce0.9Gd0.1O1.95−x for pO2 > 10−6 atm, by as much as three

b1469_Ch-12.indd 666 4/8/2013 12:40:58 PM


b1469 Catalysis by Ceria and Related Materials

Ceria and its Use in Solid Oxide Cells and Oxygen Membranes 667

orders of magnitude,272 rendering the material a MIEC under these


conditions.
All the Onsager coefficients of transport, including the cross
coefficients, were determined for Ce0.8Pr0.2O2−x within the pO2 range
10−21 to 1 atm at 800 °C.269 The cross coefficients were found to be
negligible in comparison to the direct coefficients in the pO2 range
10−21 to 10−4 atm, but of the same order of magnitude as the direct
coefficients for high pO2 values (10−2 to 1 atm).269 This observation
is in contrast with the commonly used assumption that electrons and
oxide ion vacancies migrate independently. A similar observation
has been reported for CeO2−x.273 These results provide a deeper
understanding of the mixed oxide ionic and electronic transport
properties of pure and Pr-doped ceria, and suggest that similar
investigations should be carried out for other ceria systems in order
to better understand their transport properties.

12.4.2.3 Influence of crystallite size


As discussed earlier, it is well established that oxide ion vacancies are
depleted while the number of electrons increases at the grain
boundaries of ceria, due to the presence of a positive space-charge
potential.112,158,159 Therefore the ionic conductivity of nanocrystal-
line acceptor-doped ceria is expected to be limited by the reduced
ionic conductivity across grain boundaries that lie perpendicular to
the current direction. On the other hand, the electronic conductiv-
ity is expected to be enhanced due to the contribution of highly
electronically conducting grain boundaries (due to the space-
charge potential induced increase in electronic defects) lying paral-
lel to the current direction. Tschöpe112 calculated the combinations
of grain size and acceptor concentration that define the borderline
σe = σi at 500 °C in air. A significant grain-boundary effect on the
electrolytic domain was predicted. At acceptor concentrations
exceeding 10 mol%, which are typically employed in solid electro-
lytes, the domain boundary is crossed at a grain size of ca. 20 nm.112
Recent atomistic simulations by Lee et al.274 suggest that interactions
between point defects play a major role in the dopant and oxide ion

b1469_Ch-12.indd 667 4/8/2013 12:40:58 PM


b1469 Catalysis by Ceria and Related Materials

668 C. Chatzichristodoulou et al.

vacancy concentration profiles established close to the interfaces.


According to their results pronounced segregation of Gd and V•• O
should be expected close to the grain boundaries of Gd-doped
ceria, contrary to the depletion of V•• O predicted by space-charge
theories.
Recently, Göbel et al.275 compared the conductivity of an epitaxi-
ally grown film of Ce0.9Gd0.1O1.95−x to that of two nanocrystalline
Ce0.9Gd0.1O1.95−x films with crystallite sizes of ~ 40 nm and ~ 10 nm
(film thickness was 400 nm). The single crystalline film, grown by
pulsed laser deposition (PLD) on Al2O3 (0001), showed similar
behavior to bulk microcrystalline Ce0.9Gd0.1O1.95−x. The ~ 40 nm
nanocrystalline film, grown by PLD on SiO2 (0001) at 720 °C, exhib-
ited primarily ionic conductivity in the pO2 range 10−5 to 1 atm at
280 °C, but its value was lower compared to bulk Ce0.9Gd0.1O1.95−x by
about one order of magnitude, in agreement with the expected
reduced grain-boundary ionic conductivity stemming from the
depletion of oxide ion vacancies at the space-charge layers. The
~ 10 nm nanocrystalline film, on the other hand, grown by PLD on
SiO2 (0001) at room temperature, was partly electronic conducting
under similar conditions. This is attributed to the increased grain-
boundary volume and space-charge potential in this film, resulting
in even more pronounced depletion of oxide ion vacancies and
enrichment of electrons at the grain boundaries. These observations
are in agreement with the predictions of the brick layer/space-
charge model developed by Tschöpe112 for acceptor-doped ceria.
Yeh et al.276 managed to separate the bulk and grain-boundary
contributions to the conductivity of nanocrystalline (10–100 nm)
Ce0.8Gd0.2O1.9−x in the temperature range 150–300 °C, by applying
the inverse Bonanos–Lilley equations.277 The grain core conductivity
was independent of grain size, having a value similar to that of
microcrystalline Ce0.8Gd0.2O1.9−x. The specific grain-boundary con-
ductivity was enhanced by about one order of magnitude compared
to the microcrystalline counterparts, but remained approximately
two orders of magnitude lower than the conductivity of the grain
core, resulting in a continuously decreasing total conductivity with
decreasing grain size.

b1469_Ch-12.indd 668 4/8/2013 12:40:58 PM


b1469 Catalysis by Ceria and Related Materials

Ceria and its Use in Solid Oxide Cells and Oxygen Membranes 669

A large number of investigations have focused on the fabrica-


tion and electrical characterization of nanocrystalline Gd-doped
ceria thin films during the last decade, due to their technological
importance as electrolytes in µ-SOFCs. The temperature depend-
ence of the electrical conductivity (total in-plane unless other-
wise specified in Table 12.2) under oxidizing conditions for a
large number of thin films reported in the literature is shown in
Fig. 12.13 and compared to microcrystalline Ce0.9Gd0.1O1.95−x.
Some important characteristics of these films are summarized in
Table 12.2. As shown in Fig. 12.13, most nanocrystalline thin films
have reduced conductivity and increased activation energy com-
pared to microcrystalline Ce0.9Gd0.1O1.95−x, in line with expecta-
tions from the contribution of space-charge layers at grain
boundaries. Pronounced differences in the magnitude of the
electrical conductivity are observed, though, for the various thin
films, spanning over three orders of magnitude at high tempera-
tures (~ 900 °C) and over more than five orders of magnitude at
lower temperatures (~ 200 °C).
A number of investigations have focused on the influence of
grain size on the transport properties. A decrease in activation
energy and increase in conductivity with decreasing grain size has
been generally observed,160,279 although the origin of the decreasing
activation energy remains unclear.
The pO2 dependence of the thin-film conductivity has been
investigated in some cases as a means of revealing the nature of the
observed conductivity. Measured conductivities as a function of pO2
at 600 °C are shown in Fig. 12.14. A constant conductivity value
(albeit spanning over more than two orders of magnitude for the
various samples) is observed in all cases down to a pO2 value of 10−18
atm or lower. This is generally taken as proof that ionic conductivity
prevails under these conditions. The n-type branch at lower pO2
appears to be shifted towards higher pO2 values in most of the
nanocrystalline specimens by up to three orders of magnitude com-
pared to microcrystalline Ce0.9Gd0.1O1.95−x. Both the enhanced elec-
tronic conductivity at low pO2 as well as the reduced ionic
conductivity of the nanocrystalline specimens result in a shift of the

b1469_Ch-12.indd 669 4/8/2013 12:40:58 PM


b1469_Ch-12.indd 670

670
Table 12.2 Microstructural properties of thin films included in Fig. 12.13.

Technique Gd mol% Substrate Thickness (nm) Grain size (nm) Characteristics #, Ref.
SP 20 Sapphire, 310 °C — 29 Small pores 1a,160

SP 20 Sapphire, 310 °C — 59 Small pores 1b,160


SP 20 Sapphire, 310 °C — 76 Small pores 1c,160

b1469
PLD 20 Sapphire, 300 °C — 46 Dense 2a,160
PLD 20 Sapphire, 300 °C — 55 Dense 2b,160

Catalysis by Ceria and Related Materials


C. Chatzichristodoulou et al.
PLD 20 Sapphire, 300 °C — 75 Dense 2c,160
PLD 10 MgO(100), 400 °C 20 <18 Dense, columnar, 3a,278
(111) textured
PLD 10 MgO(100), 400 °C 435 <64 Dense, columnar, 3b,278
(200) textured
SC 20 Sapphire 300–400 9 Dense 4a,279
SC 20 Sapphire 300–400 11 Dense 4b,279
SC 20 Sapphire 300–400 15 Dense 4c,279
SC 20 Sapphire 300–400 20 Dense 4d,279
SC 20 Sapphire 300–400 36 Dense 4e,279
Sputtering 10 Al2O3, 600 °C 130 50 Dense, (111) textured 5,280
EBE-IBAD Sm, 24 Al2O3, 500 °C 300–600 20 Dense 6,161
Sputtering 20 Al2O3, 300 °C 300 80 Dense, (111) textured 7,281

(Continued )
4/8/2013 12:40:58 PM
b1469_Ch-12.indd 671

Ceria and its Use in Solid Oxide Cells and Oxygen Membranes
Table 12.2 (Continued )

Technique Gd mol% Substrate Thickness (nm) Grain size (nm) Characteristics #, Ref.

b1469
Sputtering 20 Quartz 100 130 Dense 8a,282
Sputtering 20 Al2O3 100 130 Discontinuous 8b,282

Catalysis by Ceria and Related Materials


Sputtering 20 Al2O3, 600 °C 100 130 — 8c,282
PLD 20 Sapphire(0001), 700 °C 500–1,800 40 Dense, (111) textured 9a,283
PLD 20 Pt(111), 500 °C 500–1,800 90 Dense, columnar, 9b,283
(111) and ( 200)
textured, cross
plane
Sputtering 20 Pt 20 10–50 Cross plane 10a,284
Sputtering 20 Pt 50 10–50 Cross plane 10b,284
Sputtering 20 Pt 500 10–50 Cross plane 10c,284
Sputtering 20 Pt 1,500 10–50 Cross plane 10d,284
SP: spray pyrolysis, PLD: pulsed laser deposition, SC: spin coating, EBE-IBAD: electron beam evaporation-ion beam assisted deposition.

671
4/8/2013 12:40:58 PM
b1469 Catalysis by Ceria and Related Materials

672 C. Chatzichristodoulou et al.

Figure 12.13 Temperature dependence of the electrical conductivity (total in-


plane except from 9b and 10a–d that are measured cross plane) under oxidizing
conditions of Gd-doped ceria thin films reported in the literature. The total
conductivity of microcrystalline Ce0.9Gd0.1O1.95−x204 is shown for comparison. (See
Table 12.2 for a description of the materials.)

low pO2 boundary of the electrolytic domain (σi = 0.99σe) towards


higher pO2.
All possible dependences of electrical conductivity on film
thickness have been reported. Suzuki et al.279 did not observe a
thickness dependence of the conductivity of Ce0.8Gd0.2O1.9 films
deposited on sapphire by spin coating, for thicknesses between 110
nm and 630 nm, all having a similar crystallite size of ca. 20 nm.
Rodrigo et al.278 on the other hand observed a decreasing conductiv-
ity and increasing activation energy on decreasing the thickness of
nanocrystalline Ce0.9Gd0.1O1.95 films from 435 nm to 20 nm. The
films were deposited on MgO (100) substrates by PLD and had a
columnar microstructure with 10–20 nm and 10–70 nm sized crys-
tallites being predominantly (200) and (111) oriented, for the 435
nm and 20 nm thick films, respectively. Finally, Huang et al.284

b1469_Ch-12.indd 672 4/8/2013 12:40:58 PM


b1469 Catalysis by Ceria and Related Materials

Ceria and its Use in Solid Oxide Cells and Oxygen Membranes 673

Figure 12.14 PO2 dependence of the electrical conductivity of Gd-doped ceria thin
films and pellets at 600 °C and microcrystalline Ce0.9Gd0.1O1.95−x for comparison. The
crystallite size and composition of the films are provided in the legend.

reported a conductivity increase in ultrathin films (20 nm and 50


nm thick) of nanocrystalline (10–50 nm) Ce0.8Gd0.2O1.9 of more
than one order of magnitude compared to films with a thickness of
500 nm or more. The films were deposited on Si wafers coated with
500-nm-thick silicon nitride layers and 200-nm-thick Pt layers by DC
sputtering using Ce0.8Gd0.2 alloy targets and subsequent annealing
of the deposited films in air at 650 °C for 5 h in order to oxidize
them. It should be pointed out that this is the only example in the
literature of nanocrystalline Ce0.8Gd0.2O1.9 exhibiting enhanced
conductivity compared to microcrystalline Ce0.8Gd0.2O1.9 under oxi-
dizing conditions, in contrast to a wealth of experimental results
showing the opposite (see Fig. 12.13). This very interesting result
requires further investigation in order to both verify its validity and
gain more insight into the mechanism of the conductivity
improvement.
In nanocrystalline thin films, stresses within the film that result
from lattice mismatch or thermal expansion mismatch between the
substrate and the thin film or from the deposition process can
affect the electrical conduction properties.285 Also, interfacial crys-
tallographic faults may develop as a means of accommodating the
film material on the substrate, which can lead to altered transport

b1469_Ch-12.indd 673 4/8/2013 12:40:59 PM


b1469 Catalysis by Ceria and Related Materials

674 C. Chatzichristodoulou et al.

properties. Interfacial effects become more apparent with decreas-


ing thickness and this may be the reason why Suzuki et al.279 did not
observe a thickness dependence on the conductivity, as all their
films were thicker than 110 nm. The results of Rodrigo et al.278 and
Huang et al. 284 indicate that interfacial effects are significant, hav-
ing the potential to alter the electrical conductivity by more than
one order of magnitude. The opposite trends witnessed in these
two studies illustrate the important role that the deposition process
and choice of substrate have on the microstructural and electrical
properties of the interface.
The ultrathin (20–50 nm thick) nanocrystalline (10–50 nm)
Ce0.8Gd0.2O1.9 films of Huang et al.284 are the only example of
nanocrystalline Gd-doped ceria films that show higher total con-
ductivity than microcrystalline Gd-doped ceria at low temperatures
(< 500 °C). At higher temperatures, though, (> 500 °C), a large
number of thin films160,278,280,282,283 show similar or higher conductiv-
ity than microcrystalline Ce0.9Gd0.1O1.95, as shown in Fig. 12.13. This
is due to the higher activation energy generally observed for
nanocrystalline Gd-doped ceria films.
The examples given above show that the electrical properties
are crucially dependent on the microstructural parameters, such as
the degree of crystallinity, grain size, strain within the grains, strain
at the interface, level of dopant segregation, and space charge,
among others. These may vary with thermochemical conditioning,
thereby affecting the electrical properties, as the work of Karthikeyan
et al.286 has shown. Nanoscale Gd-doped ceria thin films with more
than one order of magnitude difference in electrical properties can
be obtained by suitable treatment. Furthermore, the conductivity
modulations showed good reversibility after thermochemical
conditioning.
Additional conduction mechanisms appear in doped ceria at
low temperatures (< 200 °C). Lee et al.287 investigated the reduction
and oxidation of Ce0.8Gd0.2O1.9 using conductive tip atomic force
microscopy (AFM). Their work indicated that the surface diffusion
of oxide ion vacancies dominates ionic transport in Ce0.8Gd0.2O1.9 at
temperatures up to at least 200 °C. They estimated the activation

b1469_Ch-12.indd 674 4/8/2013 12:40:59 PM


b1469 Catalysis by Ceria and Related Materials

Ceria and its Use in Solid Oxide Cells and Oxygen Membranes 675

energy of 0.56 eV for this process. Furthermore, proton conductivity


has been observed in nanocrystalline pure and Gd-doped ceria at
low temperatures (< 200 °C).288

12.5 Use of Ceria as an Electrolyte in Solid Oxide


Fuel Cells
12.5.1 Solid oxide fuel cells
A fuel cell is a device capable of converting chemical energy bound
in a fuel to electrical energy. A large number of reviews of solid
oxide fuel cells (SOFCs) can be found in the literature.289–296 The
operating principle of an SOFC is illustrated in Fig. 12.15. A solid
electrolyte, capable of conducting oxide ions, is sandwiched between
two porous electrodes (the cathode and anode). An oxidant (typi-
cally the O2 in air) is reduced at the cathode and fuel (e.g. H2, CO,
or NH3) is oxidized at the anode. More specifically, oxygen is dissoci-
ated and converted to oxide ions at the cathode/electrolyte inter-
face, whereas the electrochemical oxidation of the fuel takes place

Support
(porous)

200-1000 µm

Anode
(porous) 1-30 µm

Electrolyte
(dense) 1-30 µm

Cathode
(porous) 1-50 µm

Figure 12.15 SOFC showing the basic operational principle when the cell is oper-
ated with hydrogen and air (oxygen).

b1469_Ch-12.indd 675 4/8/2013 12:40:59 PM


b1469 Catalysis by Ceria and Related Materials

676 C. Chatzichristodoulou et al.

at the anode/electrolyte interface. For high-performance cells at a


low operating temperature, the electrolyte thickness is reduced to
1–30 µm in order to lower the corresponding ohmic resistance asso-
ciated with the conduction of oxide ions. Similarly, the anode and
cathode also have thicknesses in the 1–50 µm range. As a ceramic
structure this thin is very brittle, a support structure with good
mechanical properties is required. This support structure is typically
placed on the anode side.
The main advantage of SOFCs above a number of other electric-
ity producing technologies is their very high efficiency in converting
chemical to electrical energy. SOFCs are, in contrast to conventional
power generators such as heat engines, not bound by the thermody-
namic limit of a Carnot cycle. For a pressurized SOFC combined
with a gas turbine, efficiencies as high as 65–75% are expected.293,297
SOFCs also offer other advantages over conventional power-generat-
ing systems such as modularity (single cells can be stacked into big-
ger modules) and fuel flexibility.1,293,298 In addition, SOFCs are
almost silent and they can rapidly change their power output, so
that they can be readily incorporated into an intelligent energy
system.
SOFCs typically operate in the range 700–1000 °C and utilize an
electrolyte consisting of yttria-stabilized zirconia (YSZ). The magni-
tude of the ionic conductivity of YSZ is, however, at lower tempera-
tures inadequate for use as an electrolyte. A low operating
temperature of approximately 650 °C is considered advantageous
from a technological point of view as it would allow the use of low-
cost metallic materials such as ferritic stainless steels as intercon-
nects and supports. Lowering the operating temperature has
additional advantages, such as a more rapid start-up and a simpli-
fied thermal management system. It is also anticipated that the
degradation mechanisms that occur in SOFCs at high temperature
will be diminished. The high operating temperature of an SOFC
permits the use of a variety of fuels, such as natural gas, dimethyl
ether (DME), ammonia, methanol, and gasoline. The disadvantage
of SOFCs is their complex design, the brittle character of the cells,
and manufacturing that requires an interdisciplinary approach.

b1469_Ch-12.indd 676 4/8/2013 12:40:59 PM


b1469 Catalysis by Ceria and Related Materials

Ceria and its Use in Solid Oxide Cells and Oxygen Membranes 677

The current small scale of production, complex design, and expen-


sive materials used often leads to overall high production costs.

12.5.1.1 General requirements for electrolytes


An electrolyte material for SOFCs must possess high ionic conduc-
tivity and very low electronic conductivity at the operating condi-
tions. It must be stable in both oxidizing and reducing environments
and should ideally also be gastight. As the electrolyte has interfaces
with the anode and cathode, the electrolyte material must also be
chemically compatible with the electrode materials. An acceptable
level of ohmic resistance associated with the transport of oxide ions
through the electrolyte is ca. 0.1 Ω⋅cm2. With YSZ as the oxide ion
conductor and an electrolyte thickness of 10 µm this translates into
a minimum operating temperature of 700 °C. Using conventional,
cheap and upscalable ceramic processes it is difficult to reproduci-
bly manufacture gastight ceramic components with a thickness
below 10 µm, so materials with a higher oxide ionic conductivity
than YSZ are required in order to operate at temperatures below
700 °C. As acceptor-doped cerias possess oxide ionic conductivities
approximately one order of magnitude higher than that of YSZ at
600 °C, it is very interesting to investigate their applicability for inter-
mediate temperature operation. The above mentioned 0.1 Ω⋅cm2
and 10 µm translates into a minimum operating temperature of
500 °C for some acceptor-doped cerias.
Ceria-based materials can be used for most of the components in
an SOFC including the anode, cathode, and electrolyte. Doped ceria
has been proposed as a low-temperature electrolyte for almost half a
century7,83,193,299 and numerous studies have subsequently modeled
SOFCs with doped ceria as electrolyte, such as single cells where the
electrolyte consists of Ce1−yGdyO2−y/2−x or Ce1−ySmyO2−y/2−x. Ceria-
based materials have (in contrast to most of the zirconia-based mate-
rials) significant electronic conductivity under reducing conditions,
as already discussed in Section 12.4. This leads to a significant elec-
tronic conductivity over a large part of the electrolyte extending
from the anode side.300 Reviews on ceria as an electrolyte have been

b1469_Ch-12.indd 677 4/8/2013 12:40:59 PM


b1469 Catalysis by Ceria and Related Materials

678 C. Chatzichristodoulou et al.

carried out by Steele;216,301 however, an updated review is clearly nec-


essary as much work has been published since.
In addition to the general requirements for electrolytes men-
tioned above, the impurity level in a ceria-based electrolyte must
also be kept low, especially Si,216 as this can increase the grain-
boundary resistivity (see Section 12.4.2.1). Ceria-based electrolytes
have a significant advantage over zirconia and (La,Sr)(Ga,Mg)O3
based ones as ceria is chemically compatible with cobalt- and iron-
containing perovskites, such as (La,Sr)(Co,Fe)O3−x, which are used
as cathode materials. Cathodes made of (La, Sr) (Co, Fe)O3−x-based
materials in general have a significantly higher electrochemical
activity for oxygen reduction than cathodes made of materials that
are directly compatible with YSZ, typically (La, Sr)MnO3+/−x. The
grains in the electrolyte should preferentially be large, since this will
lower the grain-boundary resistance.
Composite electrolytes that use ceria as a backbone structure and
typically carbonates as the second phase have recently been developed.
In this review only single-phase ceria-based electrolytes will be consid-
ered even though the topic of composite electrolytes is very interesting
as very high-power densities have been demonstrated.302–305
Figure 12.16 shows the most commonly investigated geometries
of SOFCs and these are briefly discussed below. In addition to the
designs addressed below, a number of more specialized geometries

interconnect

Anode Cathodeinterconnect

Electrolyte
Electrolyte
Cathode
Cathode

Fuelflow Air flow

Air + Fuel
Flow

Figure 12.16 Different fuel cell designs. (a) Dual-chamber planar. (b) Tubular.
(c) Single-chamber planar fuel cell.

b1469_Ch-12.indd 678 4/8/2013 12:40:59 PM


b1469 Catalysis by Ceria and Related Materials

Ceria and its Use in Solid Oxide Cells and Oxygen Membranes 679

also exist.306–309 There exist also special designs for SOFCs where
the targeted power output is in the W-range and are generally
referred to as µ-SOFCs. These devices operate at relatively low tem-
peratures and are fabricated using processes mainly used in the
semiconductor industry. A recent review has been carried out by
Evans et al.310

12.5.1.2 Planar dual-chamber fuel cells


In a planar (dual-chamber) SOFC, the components are flat plates
and the fuel and oxidant (air) is separated by the electrolyte. This is
the most common design as it is easy to manufacture both in terms
of lab-scale components and in an upscaled production. A strong
advantage of this design is the short current path for electrons
between the individual cells. One disadvantage of the planar design
is that any difference in thermal expansion of the different compo-
nents and an uneven temperature distribution results in internal
stresses. This internal stress can eventually cause mechanical failure
of the cell. A recent trend in planar SOFCs has been to replace the
commonly used Ni–YSZ cermaic–metallic composite (cermet) sup-
port with a metal support.311 The metal support is mechanically
more robust than Ni–YSZ cermets. In addition, the ductility and
high thermal conductivity are beneficial for SOFCs. The thermal
expansion of steel can be tailored to that of YSZ or ceria-based elec-
trolytes and the cost of the commercial steels that can be used is
significantly lower than for the typically used cermet of nickel and
zirconia. A disadvantage of metal-supported SOFCs is the signifi-
cantly increased complexity in the manufacturing of the cells.
In addition, corrosion of the metal support could constitute a prob-
lem for long-term operation.

12.5.1.3 Tubular fuel cells


A tubular SOFC has tube geometry. Both short tubes (< 10 cm) with
a small diameter (< 5 mm) and long (> 30 cm) tubes with a large
diameter (> 10 mm) have been developed. The cell components are

b1469_Ch-12.indd 679 4/8/2013 12:41:00 PM


b1469 Catalysis by Ceria and Related Materials

680 C. Chatzichristodoulou et al.

typically deposited onto a porous electron-conducting support struc-


ture. Typically, air flows through the center of the tube and the fuel
flows around the outside of the tube (see Fig. 12.15(b)). Tubular
fuel cells have some advantages compared to planar cells, such as
better robustness during thermal cycling and under rapid changes
in operating temperature. In addition, the sealing of tubular SOFCs
is significantly easier than for planar fuel cells. Disadvantages are,
among others, more difficult production both for lab-scale compo-
nents and in large-scale manufacture. Also, some tubular designs are
hindered by a long path for the electronic current causing an addi-
tional ohmic resistance.

12.5.1.4 Single-chamber fuel cells


In a single-chamber (SC) fuel cell the anode and cathode are
exposed to the same gas (Fig. 12.16(c)). This gas is typically a non-
equilibrium mixture of air and a lightweight hydrocarbon (meth-
ane, propane, butane, or liquified petroleum gas). The cathode in
an SC-SOFC ideally only reduces oxygen and does not catalyze the
burning of the hydrocarbon present. The anode ideally only cata-
lyzes the partial oxidation of the hydrocarbon to synthesis gas (syn-
gas, a mixture of H2 and CO) as well as the oxidation of these
products. The efficiency of an SC-SOFC therefore crucially depends
on the selectivity of the electrodes. Also, a relatively low tempera-
ture must be used, typically below 600 °C, as the fuel will otherwise
be oxidized in nonelectrochemical reactions. SC fuel cells are
mainly used for portable applications where the required energy
demand is small. Compared to dual-chamber SOFCs, SC-SOFCs
have some advantages: sealing is not required, quick start-up time,
simple design and gas management, and SC-SOFCs are also
believed to be more shock resistant than conventional fuel cells
both thermally and mechanically.312–314 The main disadvantage of
the present generation of SC-SOFCs is their low efficiency and fuel
utilization due to parasitic nonelectrochemical reactions. In addi-
tion, there is a risk of explosion for fuel-air mixtures at high
temperatures.

b1469_Ch-12.indd 680 4/8/2013 12:41:00 PM


b1469 Catalysis by Ceria and Related Materials

Ceria and its Use in Solid Oxide Cells and Oxygen Membranes 681

12.5.2 Theoretical analysis


12.5.2.1 Dual-chamber planar SOFC modeling with doped
ceria as electrolyte
A number of works have addressed the modeling of ceria-based elec-
trolytes or more generally the modeling of electrolytes that are
mixed ionic and electronic conductors (MIECs). MIEC electrolytes
connected to an external electrical circuit but with reversible elec-
trodes have been discussed extensively in the literature and a few
references are given here.256,315–321 The current/voltage relations for
an SOFC with a MIEC electrolyte and irreversible electrodes have
also been the subject of a significant number of publica-
tions.216,300,319,322–328 Current research is focused on optimizing the
models with improved material parameters, better gas transport and
conversion328 or changes in design.329
A short summary of the work by Näfe,320 Dalslet et al.,325 and
Chatzichristodoulou et al.328 is given below. In principle, this work
assumes local electro-neutrality and local equilibrium.
Figure 12.17(a) shows an equivalent circuit for a single SOFC.
The electromotive force (EMF) is generated by the thermodynamic
driving force for the overall reaction. Vely is the voltage across the cell.
Iext, Ileak, and Iionic are the currents through the external circuit, the
electronic leak current, and the total current carried by oxide ions
through the electrolyte, respectively. From Kirchoff’s first law it is
clear that I ionic = I leak + I ext . For Vely < EMF (fuel cell operation)
the external and ionic current densities are considered positive
whereas the electronic current is taken to be negative. The external
current density, iext = Iext /A, where A is the electrolyte area, is given by
Vely
iext = (12.35)
AR ext

In Fig. 12.17(a), Ran and Rcat are the anode and cathode electro-
chemical polarization resistances, respectively. Rext is the external
load resistor (see below) and Rion and Relec are the resistances associ-
ated with ionic and electronic transport through the electrolyte,

b1469_Ch-12.indd 681 4/8/2013 12:41:00 PM


b1469 Catalysis by Ceria and Related Materials

682 C. Chatzichristodoulou et al.

0
839 mV (OCV)
500 mV
Iionic -5 10 m CGO10, 600 °C

log(pO / atm )
50% H 2 - 50% H2O
-10

2
-15

-20

-25

0.0 0.2 0.4 0.6 0.8 1.0


x/L
(a) (b)

Figure 12.17 (a) Equivalent circuit for an SOFC with ceria as electrolyte.
(b) Oxygen activity profile through a Ce0.9Gd0.1O1.95−x (CGO10) electrolyte with
irreversible electrodes at open-circuit voltage (OCV) and at a cell voltage of 500 mV.

respectively. Additional resistances exist, including contact resist-


ances between the different components, current constriction effects,
and diffusion for both the cathode and anode. From Fig. 12.17(a), it
can be seen that the resistance associated with the electronic leak is
connected in parallel to the external circuit, meaning that most of
the electronic current will flow through the smallest resistor.
The external resistance Rext can be expressed in terms of an
apparent internal electronic conductivity σext as R ext = s 1ext LA , where
L is the electrolyte thickness. One can then write the internal elec-
tronic current density as

RT ln(aO2 ,s'') s es i
i leak = Ú
4 FL ln(aO2 ,s ') s i + s e + s ext
d ln aO2 ( ) (12.36)

aO2 is the oxygen activity, aO2 ,s′ and aO2 ,s ″ are the oxygen activities
for the electrolyte/cathode and electrolyte/anode interfaces,
respectively. σi and σe are the ionic and electronic conductivity
of the MIEC electrolyte, respectively. The oxide ion current density
iionic can be calculated from:

RT ln(aO2 ,s'') s i (s e + s ext )


iionic = -
4 FL Úln(aO2 ,s ') s i + s e + s ext
d ln aO2 ( ) (12.37)

b1469_Ch-12.indd 682 4/8/2013 12:41:00 PM


b1469 Catalysis by Ceria and Related Materials

Ceria and its Use in Solid Oxide Cells and Oxygen Membranes 683

From the conservation of charge, the external current density can


be written as

RT ln(aΟ2 ,σ '') s εξτs ι


iext = i ionic + i leak = -
4 FL Úln(aΟ2 ,σ ') s ι + s ε + s εξτ
( )
d ln aΟ2 (12.38)

This approach corresponds to treating the external circuit as a


“virtual” electronic specie in the MIEC in addition to its intrinsic
ionic and electronic species. The oxygen activities used as integra-
tion limits in Eqs. (12.36)–(12.38) should be corrected for with the
electrode polarization losses. This is typically done using Butler–
Volmer326,327,330 or linear Chang–Jaffe216,324,325,328,331 electrode kinetics.
The linear kinetics represents a simple way to ascribe a loss of
“driving force” to each electrode as the overpotential is proportional
to the current. The oxygen activities in the integration limits of
Eqs. (12.36)–(12.38) can be calculated from:

RT Ê aO2 '' ˆ
iionic ran = ln Á ˜ (12.39)
4 F Ë aO2 ,s '' ¯

RT Ê aO2 ' ˆ
iionic rcat = - ln Á ˜ (12.40)
4 F Ë aO2 ,s ' ¯

where aO2 ′′ and aO2 ′ are the oxygen activities of the anode and cath-
ode gas, respectively, and ran and rcat are the area specific resistances
(ASRs) of the anode and cathode, respectively. These values can be
measured individually using symmetric cell measurements.
In addition, gas transport in both the support structures and the
electrodes, and variations in gas composition in the fuel flow direc-
tion should also be taken into account when modeling full
cells.326,328,329 The above represent the typical equations used for
calculating the current densities through an MIEC electrolyte.
Oxygen activity through an electrolyte at steady state can be calcu-
lated from charge conservation — ∑ J ionic = — ∑ J elec = 0 , where Jionic and
Jelec are the ionic and electronic fluxes, respectively. For the 1D case,
conservation of charge translates into ∇Jionic = ∇Jelec = 0.

b1469_Ch-12.indd 683 4/8/2013 12:41:00 PM


b1469 Catalysis by Ceria and Related Materials

684 C. Chatzichristodoulou et al.

Figure 12.17(b) shows an example of the calculated oxygen


activity through a 10-µm-thick electrolyte of Ce0.9Gd0.1O1.95−x at
600 °C as a function of the normalized thickness. The anode is
located at x/L = 0 and the cathode at x/L = 1. The calculation has
been made using material constants and polarization resistances
given by Chatzichristodoulou et al.328 The hydrogen to steam ratio
is one. Two curves are shown, one where the cell is at open-circuit
voltage (OCV) (Rext → ∞) and one where the cell is under current
(Iext = 0.66 A.cm−2 and Pext = 0.33 W.cm−2) at a cell voltage of 500
mV. The circles represent the oxygen activity in the gas phase
(inlet) and the lines the steady-state oxygen activity across the elec-
trolyte next to the inlet. The step change in oxygen activity at the
electrolyte/electrode interfaces is a representation of the overpo-
tential at the electrodes. It can be seen that the overvoltage at the
cathode/electrolyte interface is smaller than at the corresponding
anode interface, reflecting the higher anode polarization resist-
ance than for the cathode, for the chosen set of parameters. At
OCV, the step is established due to the leak current. When current
is flowing in the external circuit, the oxygen activity at the anode/
electrolyte interface is significantly larger corresponding to a
larger overpotential at the anode (more current flowing). In
effect, this means that the reduction of Ce4+ to Ce3+ is hindered,
resulting in a lower overall electronic conductivity through the
electrolyte as compared to the OCV case. Dalslet et al.325 refers to
this as the anode shielding effect. For a perfect electrolyte (no elec-
tronic conductivity) one would observe a linear relation between
the logarithm of the oxygen activity as a function of distance. Both
at OCV and under current it can be seen that oxygen activity does
not vary significantly from the anode/electrolyte interface until
approximately x/L ≈ 0.9.
From the modeling studies of SOFCs with ceria-based electro-
lytes some general conclusions can be drawn. The electronic leak
current can be very significant and must be included when mode-
ling SOFCs with a ceria-based electrolyte. Decreasing the operating
temperature strongly decreases the leak current through the elec-
trolyte as the electronic conductivity decreases more than the ionic.

b1469_Ch-12.indd 684 4/8/2013 12:41:00 PM


b1469 Catalysis by Ceria and Related Materials

Ceria and its Use in Solid Oxide Cells and Oxygen Membranes 685

Loss of performance due to electronic leakage through ceria-based


electrolytes takes place mainly at low loads and at high tempera-
tures.300,315,325,327,328,330,332 It is important to realize that increasing the
current through the external circuit will decrease the leakage cur-
rent due to a larger overpotential at the anode and thus a sup-
pressed reduction of Ce. Ceria-based electrolytes can thus work as
almost perfect electrolytes at high current densities and low tem-
peratures. Figure 12.18(a) shows the leakage current as a function
of the external current density at four different temperatures. The
calculations are again based on material parameters from
Chatzichristodoulou et al.328 It can clearly be seen that increasing the
external current density decreases the leakage current. The maxi-
mum in power output at 650 °C is Pmax = 0.61 W⋅cm−2 where iext = 1.35
A.cm−2 and the leakage current is ileak = 0.015 A⋅cm−2. At 550 °C the
maximum power output of Pmax = 0.16 W⋅cm−2 is observed at iext =
0.33 A⋅cm−2 and the leakage current at Pmax is ileak = 0.0036 A⋅cm−2.
This illustrates that a ceria-based electrolyte almost behaves as a per-
fect electrolyte (not electron conducting) at high current
densities.

0.5 1.00 0.8

10 m CGO10
0.4
50% H 2 - 50% H2O 0.75 0.6
ileak (A cm )
-2

0.3
P (W cm )

550 °C
-2

650 °C
0.50 0.4
η

0.2
650°C
10 m CGO10
0.25 10 m CGO10 ideal 0.2
0.1 600°C 97% H 2 - 3% H 2O
550°C 85% FU at Pmax
500°C
0.0 0.00 0.0
0.0 0.5 1.0
-2 -2
iext (A cm ) iext (A cm )
(a) (b)

Figure 12.18 (a) Leakage current density ileak through a 10-µm-thick Ce0.9Gd0.1O1.95−x
electrolyte as a function of the current density drawn in the external circuit iext at
500 °C, 550 °C, 600 °C, and 650 °C. (b) Efficiency η (left axis) as a function of the
external current density. The right axis shows the corresponding power density dis-
sipated in the external resistor. The fuel utilization is fixed at 85% at the maximum
power density for each temperature.

b1469_Ch-12.indd 685 4/8/2013 12:41:00 PM


b1469 Catalysis by Ceria and Related Materials

686 C. Chatzichristodoulou et al.

The (electrolyte or cell) efficiency η of a fuel cell is given


by:325,333,334
V ely iext
h= (12.41)
EMF i ionic

For a perfect electrolyte the term iext /iionic will equal one. The effi-
ciency will thus be one for a cell with a perfect electrolyte operating
at the Nernst potential. In practical terms the efficiency defined in
(12.41) is cell efficiency and does not take into account thermody-
namic efficiency nor losses due to fuel gas reformation, incomplete
fuel exploitation, and the auxiliary system.325 Figure 12.18(b) shows
cell efficiency (left axis) and the corresponding power density (right
axis) as a function of the external current density. Calculations at
650 °C and 550 °C were carried out for a 10-µm Ce0.9Gd0.1O1.95−x elec-
trolyte with the electronic conductivity of this material taken from
Chatzichristodoulou and Hendriksen272 and one where the elec-
tronic conductivity has been set to zero (which will be referred to as
a “perfect” Ce0.9Gd0.1O1.95−x electrolyte). The calculations were car-
ried out such that at the maximum power density the fuel utilization
was 85% for each temperature. Under open-circuit conditions the
perfect Ce0.9Gd0.1O1.95−x electrolyte has an efficiency of 1 whereas the
cell with the real Ce0.9Gd0.1O1.95−x electrolyte has an efficiency of zero
due to the leak current. It can be seen that the power output of the
two cases is very much alike; however, the efficiency of the SOFC
with the real Ce0.9Gd0.1O1.95−x electrolyte is much lower at low current
densities. At 550 °C and 650 °C one must exceed current densities of
approximately 0.3 A⋅cm−2 and 1.3 A⋅cm−2, respectively, in order to
achieve similar electrolyte efficiencies. However, the cell voltages at
these conditions are approximately 550 mV (1,144 mV) and 470 mV
(1,127 mV) at 550 °C and 650 °C, respectively, where the numbers in
parenthesis are the EMF values. It is thus clear that the efficiencies
of cells with a “perfect” and “real” Ce0.9Gd0.1O1.95−x electrolyte may be
similar; however this takes place at very low cell voltages and thus
very low efficiencies. The highest power densities are found at high
temperatures but the highest efficiencies are found at low

b1469_Ch-12.indd 686 4/8/2013 12:41:00 PM


b1469 Catalysis by Ceria and Related Materials

Ceria and its Use in Solid Oxide Cells and Oxygen Membranes 687

temperatures. The use of ceria electrolytes in SOFCs is therefore


clearly a matter of the required efficiency and power output. If high
power density is required with low efficiency then a ceria-based elec-
trolyte is a possibility. If high efficiency on the other hand is
required, a ceria-based electrolyte is not a good solution due to the
large leakage current.
In another example Leah et al.,332 investigating cells from
Ceres Power Ltd., found that thin-film ceria-based cells operating
at 600 °C and a typical current density of 300 mA⋅cm−2 had a leak-
age current corresponding to 6.3% of the total current. In the
study of Dalslet et al.325 several parameter variations were carried
out including variations in the anode and cathode polarization
resistances and thickness variations. Understanding the trends in
efficiency and power output is difficult due to the complex inter-
action between oxygen nonstoichiometry, the concentration of
electronic defect species, and the electrode polarization resist-
ances. For example, decreasing the anode polarization resistance
will increase the power output of a cell; however, the Ce0.9Gd0.1O1.95−x
electrolyte will become more reduced whereby the internal elec-
tronic shorting increases, thereby decreasing the overall effi-
ciency. Increasing the anode polarization resistance ensures that
the Ce0.9Gd0.1O1.95−x electrolyte is more oxidized, thereby decreas-
ing the electronic leak current. An optimum in anode polariza-
tion resistance therefore exists that will depend not only on the
operating conditions but also on the entire system. Several stud-
ies have addressed the optimal thickness and the expected OCV
for ceria-based fuel cells. Riess et al.322 and Duncan et al.327 also
found that it is not advantageous in terms of power output and
efficiency to reduce the thickness of a ceria-based electrolyte
below a certain limit in contrast to perfect electrolytes. The theo-
retical maximum power output was found for ceria-based electro-
lytes with a thickness of 1–4 µm depending on the electrodes
used. The OCV of an SOFC with a ceria-based electrolyte will also
depend on the thickness of the electrolyte and the electrodes
used, as both the electronic and ionic current densities are
nonzero. This is in contrast to an SOFC where the electrolyte is

b1469_Ch-12.indd 687 4/8/2013 12:41:01 PM


b1469 Catalysis by Ceria and Related Materials

688 C. Chatzichristodoulou et al.

zirconia-based and the leak current is negligible. A significant


scatter is thus expected in the OCV of ceria-based cells due to
differing electrode performance.330 Figure 12.19 depicts a calcu-
lation showing the maximum power output (left axis) and OCV
(right axis) for a 30-µm-thick Ce0.9Gd0.1O1.95−x membrane at 600 °C
as a function of anode and cathode area specific resistances. The
factor f is defined as f = rcat/rcat
0
and f = ran/ran
0 0
, where rcat 0
and ran are
the base values of the area specific resistances of the cathode and
anode, respectively. When the cathode area specific resistance is
varied the anode area specific resistance has the base (constant)
0
value of ran and vice versa. Conversion has not been taken into
consideration in the calculations. It can be seen that without
modification of the area specific resistance (cross over for both
curves) Pmax = 0.32 W⋅cm−2 and Vely = 0.90 V. By increasing the area
specific resistance of both the anode and cathode the OCV
decreases; however, for the cathode variation the decrease is
much stronger. Decreasing both the anode and cathode area spe-
cific resistances increases the maximum power output, as
expected, but Pmax is more sensitive to anode polarization.
However, even though the power output changes significantly,
the OCV does not change significantly but remains constant
slightly above 900 mV. The above results partly reflect that the
anode area specific resistance is significantly larger than that of
the cathode at f = 1 and at 600 °C.
It is clear that establishing a unique correlation between the
power output and OCV is not feasible; however, Fig. 12.19 illus-
trates some general tendencies. Having a relatively low power den-
sity and a high OCV could indicate that anode performance is
especially poor. Having a high power density and a high OCV could
indicate that the anode is very good (the leakage current will also
be very high). Having a low power output but a high OCV could
indicate a good cathode. Having a low power output and a low
OCV would indicate that the cathode is poor. The above tenden-
cies can serve as a guide but in principle, the OCV should be cal-
culated for every thickness and all cathode and anode area specific
resistances.

b1469_Ch-12.indd 688 4/8/2013 12:41:01 PM


b1469 Catalysis by Ceria and Related Materials

Ceria and its Use in Solid Oxide Cells and Oxygen Membranes 689

1.0
950
0.8 900

Pmax ( W cm )
-2

Vely ( mV )
0.6 850
800
0.4
750
0.2 ran rcat 700
30µm CGO10, 600 °C
50% H2 - 50% H2O 650
0.0
600
0.1 1 10
f

Figure 12.19 Maximum power and cell voltage (at OCV) as a function of the rela-
tive anode and cathode area specific resistances. f = rcat/rcat
0 0
where rcat is the base
cathode area specific resistance (rcat = 0.217 Ω⋅cm ), and similarly for the anode (r an
0 2 0

= 0.525 Ω⋅cm ). The solid and dotted lines show the power output and cell voltage,
2

respectively. The black and gray lines represent anode and cathode, respectively.
The calculation was for a 30-µm-thick Ce0.9Gd0.1O1.95−x (CGO10) electrolyte at
600 °C for a hydrogen to steam ratio of one.

12.5.2.2 Single-chamber and tubular SOFC modeling


with ceria electrolytes
A detailed study by Serincan et al.326 investigated a tubular fuel cell
with a ceria-based electrolyte. This study consisted of a furnace model
explicating transport phenomena (heat and mass flows) and consid-
ered an SOFC with a ceria-based electrolyte. Experimental data from
Suzuki et al. 335 were modeled with high precision. It was found that
at 550 °C, at a cell voltage of approximately 0.6 V, the leak currents
are negligible and the cell voltage at which leak currents become
negligible is highly temperature dependent. It was also shown that
even though the temperature in the furnace is kept at 550 °C, parts
of the cell can reach temperatures close to 668 °C at high current
densities. Axial temperature gradients as high as 18 °C⋅mm−1 were
found. In a subsequent study focusing on a transient analysis,
Serincan et al.336 found that the timescale for reaching a steady state
on changing the operating conditions for micro-tubular SOFCs was
of the order of 20 s and governed by heat transfer dynamics.

b1469_Ch-12.indd 689 4/8/2013 12:41:01 PM


b1469 Catalysis by Ceria and Related Materials

690 C. Chatzichristodoulou et al.

Few studies are available that model single-chamber SOFCs with


ceria-based electrolytes. Hao et al.337 prepared a very comprehensive
model, predicting that high power densities can be reached using
single-chamber-SOFCs with realistic material parameters. In a later
publication by Hao and Goodwin329 a comparison between a fuel
cell with a YSZ electrolyte and a Ce0.8Sm0.2O1.9 electrolyte was made.
The overall fuel cell efficiency of both types of cells is very low
(< 4%), as they are single chambers, but for temperatures below
745 °C the Ce0.8Sm0.2O1.9 electrolyte-based cell resulted in a higher
fuel-cell efficiency than the YSZ-based cell.

12.5.2.3 Stack level modeling


Only a few studies are available in the literature where a stack with
ceria-based SOFCs is modeled. An early study by Milliken et al.338
investigated a stack with a 100 mm × 100 mm × 0.25 mm ceria-based
electrolyte. For these relatively thick ceria-based electrolytes operat-
ing at 600 °C an average shorting current of 0.21 A⋅cm−2 was found at
OCV, while in operation at a cell voltage of 0.6 V the shorting current
dropped to 0.01 A⋅cm−2 while the average current density pulled
from the cell was 0.43 A.cm−2. It was found that a maximum electrical
efficiency of 41.6% was obtainable, but this decreased rapidly
as the temperature increased. In a detailed simulation study by Leah
et al.332 on two stacks with cells from Ceres Power Ltd. a maximum
predicted electrical power output of 5.1 kWe was found at a stack
efficiency of 44.3% when operating at the maximum electrical power
output. The numbers above refer to the direct current generated by
the cells and does not take into account auxiliary components, such
as air blower, control systems, or inverter. When operating the two
stacks at the maximum power output, the mean electronic leakage
was approximately 5% of the external current. When operating the
stacks at 50% of the maximum electrical power output, the stack
efficiency decreased slightly to 42.9%. For a stack containing cells
with a perfect electrolyte this number would increase due to an
increase in cell potential, but in the case of a ceria-based electrolyte
this decreases due to a higher electronic leakage. At 50% of the

b1469_Ch-12.indd 690 4/8/2013 12:41:01 PM


b1469 Catalysis by Ceria and Related Materials

Ceria and its Use in Solid Oxide Cells and Oxygen Membranes 691

maximum electrical power output the electronic leakage current was


24% of the external current. The overall energy efficiency (electrical
and heat generation) did not vary within the 50–100% electrical
power output range but remained constant at around 80%, as the
auxiliary components are utilized better at lower power outputs. The
larger electronic leakage when operating at lower power outputs is
thus transferred into more heat generation. The temperature of the
stacks was in the range 594–644 °C.
In conclusion, from a modeling point of view, ceria-based elec-
trolytes offer the clear benefit of a high oxide ion conductivity but
suffer from a high electronic leak current. In order for a cell with a
ceria-based electrolyte to be competitive against a cell with a perfect
electrolyte, the cell must be operated continuously at fairly high cur-
rent densities and at low temperatures. The specific temperature
and power density clearly depends on the given application.
However, by using ceria-based electrolytes, one of the main benefits
of SOFCs, namely the rapid upscaling and downscaling of power
output, is compromised. Nonetheless, for specific applications
where high power density is continuously required with a relatively
low stack temperature, SOFCs with a ceria-based electrolyte are
clearly advantageous.

12.5.3 Overview of performance


12.5.3.1 Dual-chamber planar SOFCs
Most of the SOFC designs with ceria-based electrolytes are planar;
however, tubular designs339–341 and single-chamber designs have also
been investigated.
Testing SOFCs with a ceria-based electrolyte is challenging
because of the leak current and this should in principle also be
measured. This can be done by exploiting the mass balance at the
fuel side of the cell. To the best of the authors’ knowledge, very few
studies have been published that address this problem,342,343 but no
studies have been published that investigate this thoroughly from an
experimental point of view.

b1469_Ch-12.indd 691 4/8/2013 12:41:01 PM


b1469 Catalysis by Ceria and Related Materials

692 C. Chatzichristodoulou et al.

Table 12.3 provides an overview of performance and Table 12.4


lists the different materials used in planar dual-chamber fuel cell tests
published in the literature, where the electrolyte is ceria based. Only
studies showing a maximum power density of more than 0.5 W⋅cm−2
at 600 °C have been included, except for a few cases where the study
is interesting in other ways. Common to all studies is that they were
carried out on anode supported cells, most commonly a cermet of Ni
and doped ceria. A few studies were carried out on metal supports or
Ni–YSZ supports. Most of the studies use the support directly as the
anode; however, a few of the studies also incorporate an anode, typi-
cally a cermet of fine-grained Ni and doped ceria. The electrolyte is,
in approximately half the studies, Gd0.1Ce0.9O1.95−x but also Sm-doped
ceria is commonly used. There is a huge variation in the composition
of the cathodes used, but in common is the use of cobalt-containing
perovskites, due to their high electrocatalytic activity towards oxygen
reduction and their thermodynamic stability with ceria-based electro-
lytes. Most of the published results were carried out on small test
areas, typically below 1 cm2. These small-scale studies focus primarily
on demonstrating a high power output, ignoring or only briefly com-
menting on the effect of leak currents.
A fairly large number of studies have demonstrated high power
outputs up to 1100 and 1329 mW⋅cm−2 at 600 °C.344,345 The OCV var-
ies unsystematically in the range 0.72–0.92 V. No correlation can be
found between the OCV of the SOFC and the power output.
The theoretical voltage at 600 °C for H2 humidified with 3% water is
1.11 V. None of the cells reached the theoretical OCV value due to
the internal electronic short circuit. The huge scatter in OCV is
consistent with the fact that it is a complex function of the thickness
of the electrolyte, electrode polarization resistances, and electrical
contact to the cell as illustrated in Fig. 12.19. Some of the studies
listed in Tables 12.3 and 12.4 are discussed in more detail below.
Ahn et al.345 investigated346 a Ce0.85Sm0.075Nd0.075O2−x electrolyte as
recent studies indicate that this composition has a 30% higher ionic
conductivity than the commonly used Gd-doped ceria.207,347 A very
high performance of 1100 mW.cm−2 was obtained at 600 °C.
Micrographs in the study of Ahn et al. show surprisingly large grains

b1469_Ch-12.indd 692 4/8/2013 12:41:01 PM


b1469_Ch-12.indd 693

Table 12.3 Performance of fuel cells with a ceria-based electrolyte. If not specified the anode gas is humidified (~3%) hydrogen
and the cathode gas is air. Further details of the design of the cells are in Table 12.4.
Pmax Pmax Pmax Pmax

Ceria and Its Use in Solid Oxide Cells and Oxygen Membranes
Area (650°C) (600°C) (550°C) (500°C) OCV OCV OCV OCV
Ref. Electrolyte cm2 mW⋅cm−2 mW⋅cm−2 mW⋅cm−2 mW⋅cm−2 (650°C) V (600°C) V (550°C) V (500°C) V

b1469
353 15 µm, Ce0.8Sm0.2O1.9 869 648 385 220 0.74 0.75 0.78 0.83
345 10 µm, Ce0.85Sm0.075 0.48 1430 1100 730 320 0.86 0.89 0.93 0.96

Catalysis by Ceria and Related Materials


Nd0.075O1.925
346 10 µm, Ce0.9Gd0.1O1.95 994 913 627 440 0.796 0.830 0.874 0.913
354 20 µm, Ce0.8Sm0.2O1.9 750 490 260 120 0.91 0.93 0.95 0.97
348 2 µm, Ce0.9Gd0.1O1.95 440 370 250 0.81 0.83 0.85
348 4 µm, Ce0.9Gd0.1O1.95 771 548 325 0.91 0.94 0.96
348 16 µm, Ce0.9Gd0.1O1.95 644 363 174 0.92 0.94 0.96
351 30 µm, Ce0.8Sm0.2O1.9 0.34 216 176 120 75 0.768 0.83 0.87 0.92
355 10 µm, Ce0.8Gd0.2O1.9 0.5 578 358 167 0.863 0.901 0.95
344 10 µm, Ce0.9Gd0.1O1.95 0.5 1329 863 454 0.903 0.984
356 30 µm, Ce0.9Gd0.1O1.95 0.5 443 231
357 22 µm, Ce0.9Gd0.1O1.95 0.5 863 562 321 154 0.843 0.884 0.916 0.952
358 20 µm, Ce0.8Sm0.2O1.9 0.5 910 750 460 230 0.81 0.85 0.88 0.89
350 15 µm, Ce0.9Gd0.1O1.95 620 380 180 0.70 0.72 0.64

(Continued )

693
4/8/2013 12:41:01 PM
b1469_Ch-12.indd 694

694
Table 12.3 (Continued )

Pmax Pmax Pmax Pmax


Area (650°C) (600°C) (550°C) (500°C) OCV OCV OCV OCV
Ref. Electrolyte cm2 mW⋅cm−2 mW⋅cm−2 mW⋅cm−2 mW⋅cm−2 (650°C) V (600°C) V (550°C) V (500°C) V

b1469
350 25 µm, Ce0.9Gd0.1O1.95 680 420 210 0.83 0.84 0.85
359 20 µm, Ce0.85Sm0.15O1.925 ~1 1,010 ~640 402 ~0.860 ~0.900

Catalysis by Ceria and Related Materials


C. Chatzichristodoulou et al.
360 40 µm, Ce0.9Gd0.1O1.95 0.5 769 0.890
750 0.890
(CH4) (CH4)
716 0.880
(C2H6) (C2H6)
648 0.870
(C3H8) (C3H8)
361 20 µm, Ce0.9Gd0.1O1.95 0.5 ~900 ~590 ~330 0.916 0.959 0.999
349 35 µm, Ce0.8Sm0.2O1.9 700 491 185 65 0.86 0.90 0.94 0.95
(BPG a) 334 (BPG a) (BPG a) 0.90
(BPG a)
362 20 µm, Ce0.9Gd0.1O1.95 0.25 694 602 432 230 0.787 0.852 0.869 0.889
342 10 µm, Ce0.8Sm0.2O1.9 1.35 595 545 347 193 0.796 0.835 0.857 0.873
Note: a BPG is biomass-produced gas (14.7% CO, 14.2% CO2, 15.3% H2, 4.2% CH4, and 51% N2).
4/8/2013 12:41:01 PM
b1469_Ch-12.indd 695

Table 12.4 Design of fuel cells in Table 12.3.

Ref. Support Anode Cathode Electrolyte


353 NiO, 65% Support Ba0.5Sr0.5Co0.8Fe0.2O3−x, 70% 15 µm, Ce0.8Sm0.2O1.9
Ce0.8Sm0.2O1.9, 35% Ce0.8Sm0.2O1.9, 30%

Ceria and its Use in Solid Oxide Cells and Oxygen Membranes
345 NiO, 65% Ce0.9Gd0.1O1.95 La0.6Sr0.4Co0.2Fe0.8O3−x, 50% 10 µm
Ce0.9Gd0.1O1.95, 35% (Anode functional Ce0.9Gd0.1O1.95, 50% Ce0.85Sm0.075Nd0.075O1.925

b1469
layer)
346 NiO, 65% Ce0.9Gd0.1O1.95 (Anode La0.6Sr0.4Co0.2Fe0.8O3−x, 50% 10 µm Ce0.9Gd0.1O1.95

Catalysis by Ceria and Related Materials


Ce0.9Gd0.1O1.95, 35% functional layer) Ce0.9Gd0.1O1.95, 50%
Variations in anode
354 Hastelloy NiO, 50% Sm0.5Sr0.5CoO3−x, 70% 20 µm, Ce0.8Sm0.2O1.9
Ce0.8Sm0.2O1.9, 50% Ce0.8Sm0.2O1.9, 30%
348 NiO Support La0.8Sr0.2Co0.8Fe0.2O3−x 1–75 µm, Ce0.9Gd0.1O1.95
Ce0.9Gd0.1O1.95
351 Hastelloy X, metal NiO, 70% Sm0.5Sr0.5CoO3−x, 75% 30 µm, Ce0.8Sm0.2O1.9
Ce0.8Sm0.2O1.9, 30% Ce0.8Sm0.2O1.9, 25%
355 NiO, 65% Support La0.6Sr0.4Co0.2Fe0.8O3−x, 50% 10 µm, Ce0.8Gd0.2O1.9
Ce0.9Gd0.1O1.95, 35% Ce0.9Gd0.1O1.95, 50%
344 NiO, 65% Support Ba0.5Sr0.5Co0.8Fe0.2O3−x 10 µm, Ce0.9Gd0.1O1.95
Ce0.9Gd0.1O1.95, 35%
356 NiO Support Bi2V0.9Cu0.1O5.35, 43% 30 µm, Ce0.9Gd0.1O1.95
Ce0.9Gd0.1O1.95 Ag, 57%
357 NiO, 70% Support La0.8Sr0.2Co0.2Fe0.8O3−x, 50% 22 µm, Ce0.9Gd0.1O1.95

695
4/8/2013 12:41:01 PM

Ce0.9Gd0.1O1.95, 30% Ce0.9Gd0.1O1.95, 50%


(Continued )
b1469_Ch-12.indd 696

696
Table 12.4 (Continued )

Ref. Support Anode Cathode Electrolyte


358 NiO, 61.3% support La0.8Sr0.2Co0.2Fe0.8O3−x 20 µm, Ce0.8Sm0.2O1.9
Ce0.8Sm0.2O1.9, 38.7%
350 Fe2O3, 30 mol% NiO, 50% La0.6Sr0.4CoO3−x, 50% 15 µm, Ce0.9Gd0.1O1.95
NiO, 70 mol% Ce0.9Gd0.1O1.95, 50% Ce0.9Gd0.1O1.95, 50% 25 µm, Ce0.9Gd0.1O1.95

b1469
359 NiO, 60% Support Sm-doped ceria porous 20 µm, Ce0.85Sm0.15O1.925
Ce0.85Sm0.15O1.925, 35% interlayer (<5 µm)

Catalysis by Ceria and Related Materials


C. Chatzichristodoulou et al.
Ba0.5Sr0.5Co0.8Fe0.2O3−x
360 NiO Support Sm0.5Sr0.5CoO3−x 40 µm, Ce0.9Gd0.1O1.95
Ce0.9Gd0.1O1.95, 50 wt%
CuO, PdO, PtO3,
Rh2O3, RuO2
361 NiO, 50% Support Ce0.9Gd0.1O1.95, interlayer 20 µm, Ce0.9Gd0.1O1.95
Ce0.9Gd0.1O1.95, 50% Ba0.5Sr0.5Co0.8Fe0.2O3−x, 70%
Ce0.9Gd0.1O1.95, 30%
349 NiO, 65% Support Sm0.5Sr0.5CoO3−x, 70% 35 µm, Ce0.8Sm0.2O1.9
Ce0.8Sm0.2O1.9, 35% Ce0.8Sm0.2O1.9, 30%
362 NiO, 65% Support Sm0.5Sr0.5CoO3−x, 70% 20 µm, Ce0.9Gd0.1O1.95
Ce0.9Gd0.1O1.95, 35% Ce0.9Gd0.1O1.95, 30%
342 NiO, 57% NiO, 53% Sm0.5Sr0.5CoO3−x, 75% 10 µm, Ce0.8Sm0.2O1.9
Zr0.84Y0.16O1.92, 43% Ce0.8Sm0.2O1.9, 47% Ce0.8Sm0.2O1.9, 25%
4/8/2013 12:41:01 PM
b1469 Catalysis by Ceria and Related Materials

Ceria and its Use in Solid Oxide Cells and Oxygen Membranes 697

of the order 10–20 µm, which are very large compared to those seen
in other studies (typically 1–2 µm). In general large grains are pref-
erable as the total grain-boundary resistance will be lower.
A thorough study by Ding et al.348 investigated a series of anode
supported SOFCs with a Ce0.9Gd0.1O1.95 electrolyte and thicknesses in
the range from 1 to 75 µm. Only the cells with the best performance
(electrolyte thicknesses: 2, 4, and 16 µm) are listed in Table 12.3. It
was claimed that electrolyte thicknesses down to 1 µm could be
made gastight. The OCV increased strongly from 0.81 to 0.95 V on
changing the electrolyte thickness from 2 to 75 µm, using humidi-
fied hydrogen as fuel (3% H2O). Most of the change in OCV took
place between electrolyte thicknesses of 2 and 4 µm. This is consist-
ent with the theoretical work of Duncan et al.327 A maximum power
density of 771 mW⋅cm−2 at 600 °C was found for the 4-µm-thick elec-
trolyte. Increasing the thickness to 16 µm decreased the power den-
sity to 644 mW⋅cm−2. Decreasing the electrolyte thickness to 1 µm
strongly decreased the maximum power output to 94 mW⋅cm−2.
In a study by Yin et al.349 a fuel derived from the gasification of
rice husks was used. The composition of the fuel was CO (14.7%),
CO2 (14.2%), H2 (15.3%), CH4 (4.2%), and N2 (51%). The biomass-
produced gas (BPG) fed an SOFC with a 35-µm-thick Ce0.8Sm0.2O1.9
electrolyte. Feeding pure hydrogen to the cell resulted in a power
output of approximately 491 mW.cm−2 while the BPG resulted in a
power output of 334 mW⋅cm−2. The OCV for humidified hydrogen
(3%) and the BPG were almost equal at 0.91 V at 600 °C. However,
the theoretical equilibrium OCV of the BPG is much lower than that
of pure hydrogen. This suggests that ceria can possibly work as an
electrolyte in cells that operate with BPG as the fuel, as the theoreti-
cal OCV of the gas is lower than that found in hydrogen/methane
atmospheres. The less reducing conditions on the anode side of the
fuel cell cause a lower electronic conductance through the electro-
lyte and thus a reduced parasitic contribution.
In a detailed study by Zhang et al.342 an SOFC with a 10-µm
Ce0.8Sm0.2O1.9−x electrolyte and a Ni–Ce0.8Sm0.2O1.9−x anode of 15 µm
on a Ni–YSZ substrate was investigated. In this study the anode of
gas was collected, condensed and weighed. It was found that at

b1469_Ch-12.indd 697 4/8/2013 12:41:01 PM


b1469 Catalysis by Ceria and Related Materials

698 C. Chatzichristodoulou et al.

600 °C the leakage current at OCV was 1.06 A.cm−2. Running the cell
at 0.5 A⋅cm−2 (the cell voltage was 0.69 V) reduced the leak current
to 0.82 A⋅cm−2. In this study the leak current through a cell without
a cathode was also tested. The leak current in this case decreased to
approximately 0.3 A⋅cm−2 indicating that gas leakage through the
electrolyte is not responsible for the large leak current measured.
Park and Virkar350 investigated SOFCs with a 15-µm and a
25-µm-thick Ce0.9Gd0.1O1.95 electrolyte. The support was in all cases
a bimetallic support of Ni and Fe and the anode was a cermet of Ni
and Ce0.9Gd0.1O1.95. The best performance was found for the 25-µm-
thick Ce0.9Gd0.1O1.95 electrolyte with a power density of 0.42 W⋅cm−2
at 600 °C. Reducing the thickness of the electrolyte layer to 15 µm
also reduced the power output slightly (from 0.42 W⋅cm−2 to
0.38 W⋅cm−2); however the OCV reduced drastically to approxi-
mately 0.70 V. The decrease in OCV with decreasing thickness was
attributed to the overall lower ionic transference number for thin-
ner membranes.
Wang et al.351 investigated a metal-supported planar SOFC with a
Ce0.8Sm0.2O1.9−x electrolyte. A power density of 0.176 W⋅cm−2 was
obtained at 600 °C. In a subsequent AC study by Huang et al.352 on
similar cells it was estimated that the overall resistance was domi-
nated by contact resistance due to oxidation between the metal
support and the anode layer.

12.5.3.2 Single-chamber and tubular SOFCs


Hibino et al.312 investigated a SC-SOFC with a 0.150-mm-thick
Ce0.8Sm0.2O1.9 electrolyte. The cathode was Sm0.5Sr0.5CoO3−x while
the anode consisted of 10 wt% Ce1−ySmyO2−y/2−x-containing Ni.
Methane, ethane, and propane were used as fuels. For a compara-
tive study both YSZ and La0.9Sr0.1Ga0.8Mg0.2O3 electrolytes were also
investigated. It was found that the OCV of the three different cells
were very similar at around 0.92 V at 773 K, indicating that the elec-
tronic short circuiting in the Ce0.8Sm0.2O1.9 electrolyte is not so sig-
nificant for SC-SOFCs at these low temperatures. Impressive power

b1469_Ch-12.indd 698 4/8/2013 12:41:01 PM


b1469 Catalysis by Ceria and Related Materials

Ceria and its Use in Solid Oxide Cells and Oxygen Membranes 699

densities of 403 and 101 mW⋅cm−2 were reported at 500 °C and


400 °C, respectively.
Shao et al.313 employed a 20-µm Ce0.85Sm0.15O1.925 electrolyte on
a support of Ni and Ce0.85Sm0.15O1.925 (NiO: Ce0.85Sm0.15O1.925 =
60:40 wt%). The cathode was a composite of Ba0.5Sr0.5Co0.8Fe0.2O3−x
(70 wt%) and Ce0.85Sm0.15O1.925 (30 wt%). Due to the exothermic
nature of the reaction, the cell temperature could be as much as
150 °C higher than the furnace temperature. A peak power output of
760 mW⋅cm−2 was reached at a cell temperature of 790 °C (furnace
temperature 650 °C). The temperature of the cell was measured using
a thermocouple physically connected to the anode support. It can
clearly be seen that precise temperature measurement is a necessity
when characterizing single-chamber fuel cells, but also in ordinary
dual-chamber SOFCs, precise cell temperature measurements are
important in order to interpret the results correctly. The cell voltages
in the study of Shao et al. were fairly constant between 0.70–0.78 V.
Suzuki et al.335,340,341 demonstrated so-called micro-tubular sys-
tems with ceramic tubes consisting of a Ni/Ce0.8Gd0.2O1.9 support
structure (thickness 200 µm) with an approximate 20-µm-thick
Ce0.8Gd0.2O1.9 electrolyte. Both the cathode of La0.6Sr0.4Co0.2Fe0.8O3−x
and the electrolyte were prepared using dip coating. The diameter
was around 1.6 mm after sintering. A single tube had power densi-
ties of 400 and 1000 mW⋅cm−2 at 500 °C and 570 °C, respectively.
A disadvantage of tubular fuel cells is the resistance due to electron
transport in the support structure, which was also pointed out by
Suzuki et al.335

12.5.4 Technological programs


Much has been published on ceria-based fuel cells; however, to date,
only Ceres Power Ltd., Crawly, UK, is developing SOFCs with ceria-
based electrolytes, whereas numerous companies are developing
SOFCs with zirconia-based electrolytes. Ceres Power Ltd. is develop-
ing a ceria-based electrolyte on a metal support, where the target
operating temperature is 500–600 °C.363,364 Ceres Power Ltd. is focus-
ing on the development of fuel cells for small-scale combined heat

b1469_Ch-12.indd 699 4/8/2013 12:41:01 PM


b1469 Catalysis by Ceria and Related Materials

700 C. Chatzichristodoulou et al.

and power (CHP) products for the residential sector and in energy
security applications (http://www.cerespower.com). Ceres Power
Ltd. uses a metal substrate onto which different active layers are
deposited. Not much information is available either on processing
details or on the performance of the fuel cells. From the few publi-
cations available it is clear that the electrodes are deposited using
standard ceramic processing techniques (wet spraying and screen
printing). An electrolyte is produced using an electrophoretic depo-
sition technique and it is claimed that a sintering temperature
below 1000 °C can be used. The anode is a Ni–Ce1−yGdyO2−y/2−x cer-
met and the cathode is doped lanthanum ferrite. At 600 °C a power
density of 310 mW⋅cm−2 can be obtained with 4 × 4 cm2 cells. Ceres
Power is currently (2011) conducting field tests of their CHP units.
During these field tests problems with reliability and durability
emerged, but it was also claimed that these have been solved. Recent
presentations at international conferences indicate that Ceres
Power Ltd. will switch from producing a completely ceria-based elec-
trolyte to an electrolyte consisting of a thin zirconia-based electro-
lyte (electron blocking) and a thicker ceria electrolyte.
From a publication by Huijsmans et al.365 it is also evident that
the Netherlands Energy Research Foundation (ECN) has upscaled
planar SOFCs with a Ce0.9Gd0.1O1.95 electrolyte and that these cells
have been evaluated in the Sulzer Hexis concept.
Based on a number of publications it is clear that the National
Research Council of Canada (NRC) has a very ambitious initiative for
SOFCs with ceria-based electrolytes.342,351,354,366–369 Most of the studies
have focused on metal-supported (Hastelloy) SOFCs and a few of
them have used traditional cermet structures. It is interesting to notice
that even though the NRC clearly are aware of the internal shorting of
the cells342 a strategy for handling this has not been presented.

12.5.5 Interdiffusion barrier layer


Cathodes with a high electrochemical performance are typically
based on cobalt- and iron-containing perovskites such as (La, Sr)
(Co, Fe)O3−x.359,370–372 These cathode materials, especially those

b1469_Ch-12.indd 700 4/8/2013 12:41:01 PM


b1469 Catalysis by Ceria and Related Materials

Ceria and its Use in Solid Oxide Cells and Oxygen Membranes 701

containing Co, react with zirconia-based electrolytes forming the


insulating phases SrZrO3 and La2Zr2O7 at temperatures as low as
800 °C.28,373 As ceria-based materials are generally stable against iron-
and cobalt-based perovskites, doped ceria is also used as an interdif-
fusion barrier/buffer layer between the cathode and an electrolyte
based on zirconia. The requirements of the ceria-based interdiffu-
sion barrier layer is that (i) it must be thin in order to minimize the
additional ionic resistance, (ii) ideally it should be dense or at least
gastight in order to prevent gaseous Sr diffusion from the cathode to
the zirconia-based electrolyte,374 and (iii) if feasible the grain size
should be large in order to avoid diffusion of typically Sr through the
grain boundaries.375 Both of the requirements (ii) and (iii) suggest a
high sintering temperature for this interdiffusion barrier layer; how-
ever, as ceria and YSZ at high temperatures (> 1200 °C) form a solid
solution with low ionic conductivity this suggests a low sintering tem-
perature.203,376,377 There has been a large number of studies on the
fabrication of these layers and the testing of full cells. Typical fabrica-
tion procedures are screen printing,378–381 slurry spraying, spin coat-
ing,382 metal organic deposition (MOD),383 physical vapor deposition
(PVD),381,384 pulsed laser deposition (PLD),372 dip coating, 385–387 and
infiltration.388 Another route for obtaining dense layers is by adding
sintering aids, cobalt particularly seems very effective in this
respect.380,387,389 Barrier layers produced by physical processes such as
PVD and PLD usually tend to result in a better performance than
layers produced using conventional techniques.381,382,390

12.6 Use of Ceria for Oxygen Membranes


Due to its high ionic conductivity and good chemical stability, accep-
tor-doped ceria can potentially be used in dense ceramic mem-
branes for the separation of oxygen from air.

12.6.1 Oxygen
Oxygen is an important industrial gas annually produced in large
quantities (100 million tonnes/year).391 Typically, production takes

b1469_Ch-12.indd 701 4/8/2013 12:41:01 PM


b1469 Catalysis by Ceria and Related Materials

702 C. Chatzichristodoulou et al.

place in large cryogenic distillation plants and the gas is delivered


to an oxygen-consuming process at the same site or is distributed in
liquid or pressurized form. For specific applications, where purity
requirements (< 94%) and daily consumption (10–5000 Nm3⋅h−1)
are modest, oxygen may also be produced on-site by pressure swing
adsorption (PSA) or vacuum pressure swing adsorption (VPSA).
For large-scale use, in conjunction with e.g. ammonia and metha-
nol production, cryogenic distillation is preferred and oxygen
plants are built next to the synthesis plants. This is also the case in
some large steel and glass plants. Cryogenic distillation plants are
capital intensive and the production process is very energy consum-
ing (210 kWh⋅t−1).392

12.6.2 Gas separation membranes


Alternatively, oxygen may be produced in a separation process using
gas separation membranes, which is the topic of intense R&D393 in
both industry394,395 and academia. This research is driven by the poten-
tial of the membrane process to reduce energy consumption and the
cost of the oxygen production process. Energy savings of 30–50%392
have been estimated. Membrane technology is still at the R&D stage.
A multi-partner consortium led by Air Products is the world leader in
upscaling the membrane process. It has proved the concept in a dem-
onstration plant with a nominal production capacity of 5 tons per day
(TPD) and has announced the start of a 100 TPD plant. Other gas
supply companies also carry out research in this area.394,395
The membrane route for oxygen production relies on the use of
oxide ion conducting materials in a dense form. The oxygen is trans-
ported through the membrane in the form of ions and electrons are
transported in the opposite direction. At both surfaces of the mem-
brane an exchange process takes place between molecular gaseous
oxygen, electrons, and oxide ions. The concept is illustrated in
Fig. 12.20 for two slightly different membranes and three different
applications (described in 12.6.3 and 12.6.4).
The driving force for the oxygen flux is the difference in oxygen
activity between the two sides of the membrane. If the oxygen

b1469_Ch-12.indd 702 4/8/2013 12:41:01 PM


b1469 Catalysis by Ceria and Related Materials

Ceria and its Use in Solid Oxide Cells and Oxygen Membranes 703

Oxyfuel, “Three end” Oxyfuel,“Four end” Syngas


CH4 , H2O H2, CO, (CO2,H 2O)
Pure O2 CO2, H2O CO2, H2O, O2 PO2 ~ 10-19 atm
1 atm PO2 ~ 1 atm 1 atm
Catalyst layer PO PO2~0.2
2~
0,2atm
atm.
1 atm
e -
e- e-
2-
O O 2- e- O2-
O2- 1 atm
20 atm PO2 ~ 4 atm PO2 ~ 0.2 atm
Air O2 depleted air 10atm PO2 ~ 2 atm Air O2 depleted air
Air O2 depleted air

Figure 12.20 Oxygen permeation membranes for three different applications. In


all applications one side of the membrane is purged with air. In the oxyfuel applica-
tion the membrane may be part of a standalone unit producing pure oxygen (left)
or it may be more closely integrated with a flue gas stream, which passes over the
membrane, becomes enriched with oxygen and is recycled to a boiler (center). For
syngas, methane (and steam) is fed to the permeate side of the membrane to
become partially oxidized. The membrane may be a single-phase material showing
mixed oxide ionic and electronic conductivity (center or right) or it may be a com-
posite of an electronically conducting and an ionic conducting phase (left).

exchange process on the two surfaces is fast the flux through the
membrane is given by:

ln(aO2 ') s is e
J O2 =
RT
2 Úln(a '')
16 F L O2 si + se
d ln aO2 ( ) (12.42)

where L is the membrane thickness, T the absolute temperature,


R the gas constant, σi the ionic conductivity, and σe the electronic
conductivity. aO2′ is the oxygen activity on the high pO2 side of the
membrane and aO2″ the oxygen activity on the low pO2 side (aO2 =
pO2/1 atm). The cost of a membrane plant for oxygen production
will scale in inverse proportion to the flux. Evidently from Eq. (12.42)
to maximize the flux one needs to maximize the difference in pO2
over the membrane, minimize the membrane thickness, and use
materials with high ionic as well as electronic conductivity (the lower
of the two limits the performance).
As the membrane is dense and oxygen is transported in the
form of oxide ions, high selectivity is easily achievable with this
design. Appreciable production rates require the membrane to
operate at a high temperature (500–1000 οC) where ionic

b1469_Ch-12.indd 703 4/8/2013 12:41:01 PM


b1469 Catalysis by Ceria and Related Materials

704 C. Chatzichristodoulou et al.

conduction in typical oxide ion conductors is significant (of the


order of 10−3 to 10−1 S⋅cm−1).
Equation (12.42) describes the oxygen flux through a mixed
conducting membrane where the oxygen exchange process (molec-
ular oxygen to oxide ions in the lattice; O2 + 4e−↔ 2O− −, see
Fig. 12.20) is very fast compared to the bulk transport. This will not
always be the case. For good conductors and when considering thin
membranes the rates of the exchange process may be comparable
to the rate of diffusion or even become the overall rate limiting for
the oxygen transport. A number of different ways of taking into
account oxygen permeation through the membrane are described
in the literature.393,396–398 A simple way is to ascribe to each surface a
loss of “driving force” proportional to the current (linear kinetics)
and the area specific resistance, r, and to correct the integration
limits of Eq. (12.42) accordingly. This approach was discussed in
Section 12.5.2.1, and the new integration limits aO2,s′ and aO2,s″ to be
applied in Eq. (12.42) can be calculated directly from Eqs (12.40)
and (12.39) in Section 12.5.2.1. The area specific resistances may
be directly measured where the oxygen flux is electrically (electro-
chemically) driven, or it may be deduced from the values of the
surface exchange rate constant k0 measured in tracer diffusion
experiments:

RT 1
r = (12.43)
4 F 2 C 0 k0

where C0 is the oxygen concentration in the material right at the


surface (evaluated at aO2,s′ and aO2,s″ for the anode (permeate) and
cathode (feed) reactions, respectively). The resistance may alterna-
tively be deduced from measurements of the chemical surface
exchange rate constant, kex = γko, where γ = 1/2∂ ln pO2/∂ ln CO (for
mixed conductors with large electronic conductivity398). From the
above parameters a characteristic length may conveniently be
introduced396:

lc = D0/k0 = σir (12.44)

b1469_Ch-12.indd 704 4/8/2013 12:41:02 PM


b1469 Catalysis by Ceria and Related Materials

Ceria and its Use in Solid Oxide Cells and Oxygen Membranes 705

where D0 is the oxygen component diffusion coefficient. The char-


acteristic length is the membrane thickness where the loss of driving
force distributes equally between the surface and the bulk.
Reducing the membrane thickness will result in improved flux
(cf. Eq. (12.42)) as long as the membrane thickness is large com-
pared to lc, whereas reducing the thickness much below the charac-
teristic length is not efficient for maximizing the flux, as it becomes
primarily controlled by the rate of surface reactions.

12.6.3 Applications of oxygen gas separation membranes


Membrane can best compete with existing oxygen production tech-
nologies where thermal integration of the membrane reactors in the
plant receiving the oxygen is possible. Hence, most current research
in membranes and membrane reactors focuses on oxygen mem-
branes for specific high-temperature processes. Two important exam-
ples of such processes are oxyfuel processes in fossil power plants and
plants for production of synthesis gas by autothermal reforming of
natural gas.
In an oxyfuel process, oxygen mixed with recycled flue gas is
used in the combustion process rather than air, which has the advan-
tage that downstream CO2 capture becomes much simpler as the
flue gas contains no N2 and a separation of CO2 and N2 is thus not
required. Research into oxyfuel processes is motivated by the need
to mitigate emissions of greenhouse gases from fossil fuel power
plants. Two different schemes for the integration of the membrane
in the power plant process have been suggested: (i) direct integra-
tion (sometimes referred to as the four-end system), where one side
of the membrane is exposed to a recycle stream from the boiler,
which after oxygen enrichment is fed back to the boiler (Fig. 12.20
center) and (ii) a “standalone” (or three-end system), where a
stream of pure oxygen is produced and subsequently mixed with a
recycle stream from the boiler (Fig. 12.20 left). To maximize overall
efficiency, close thermal integration should be ensured for both
schemes. The two different integration schemes result in different
chemical environments for the membrane and thus in different

b1469_Ch-12.indd 705 4/8/2013 12:41:02 PM


b1469 Catalysis by Ceria and Related Materials

706 C. Chatzichristodoulou et al.

material requirements and in different options as to how to ensure


a large driving force across the membrane. Flue gas from a coal-fired
boiler typically contains around 2–5% oxygen, 70% N2, ~12–15%
CO2, and 10–15% H2O. Hence, using ambient air on the high pO2
side of the membrane will result in a driving force for oxygen per-
meation to the recycle side. The driving force can be increased by
pressurizing the air side of the membrane. In the “standalone” case
one needs a vacuum on the permeate side or a pressure (above
5 bar) on the air side, in order to ensure a driving force in the direc-
tion of the oxygen stream. The advantages of direct integration is
that it makes thermal integration easier and that the use of the flue
gas on the low pO2 side adds to the driving force without having an
energy penalty (which is the case if the increase in driving force is
achieved via a vacuum or increased pressure on the high pO2 side).
Using system modeling the energy efficiency penalty due to CO2
capture of an oxyfuel coal-fired power plant based on the four-end
integration scheme has been estimated to be ~5 percentage points
compared to 5–7 percentage points for the three-end scheme, which
also requires about twice the membrane area.399 The energy penalty
of the oxyfuel/membrane route for CO2 capture is estimated to
be as low as half of that of alternative post-combustion capture
routes.399,400 Direct integration on the other hand imposes more
stringent material requirements such as material stability towards
CO2 and pollutants in the flue gas like alkalis and sulfur.
Another important application targeted in membrane develop-
ment projects is the use of membranes for the production of synthe-
sis gas. Synthesis gas is an important intermediate in the production
of ammonia and methanol both of which are important chemicals
produced globally in huge amounts: in 2010 ammonia production
exceeded 130 million tonnes and methanol production exceeded
45 million tonnes.401,402 Synthesis gas is today produced in large
quantities by the autothermal reforming of natural gas where the
endothermic steam reforming process is combined with exothermal
partial oxidation in a manner that makes the overall process “ther-
moneutral.” Oxygen for the process is supplied from a cryogenic air
separation unit on site. Here, as with the oxyfuel process, one can

b1469_Ch-12.indd 706 4/8/2013 12:41:02 PM


b1469 Catalysis by Ceria and Related Materials

Ceria and its Use in Solid Oxide Cells and Oxygen Membranes 707

envisage the oxygen-producing membrane simply replacing the


cryogenic process (i.e. a standalone plant) or direct integration
where the low pO2 side of the membrane is directly exposed to the
natural gas stream or a partially reformed gas stream. The latter is
highly advantageous in terms of overall process efficiency as thermal
integration is simplified and a large driving force for the oxygen
permeation is realized (at no energy expense) by exposure of the
membrane to the natural gas/synthesis gas. The latter gas mixture
has an exit pO2 of the order of 10−19 atm at 1000 °C and hence the
materials to be used in this application must be stable under strongly
reducing conditions.
In addition, oxygen membranes have been suggested for use in
conjunction with oxygen supply/retrieval from metal melts403 and to
supply oxygen to a range of chemical processes, e.g. the upgrade of
methane to ethene or the production of hydrogen cyanide via the
Andrussow reaction.404 Other interesting applications of oxygen
membranes include biomass gasification and oxyfuel processes in
high-temperature energy-intensive industries like the steel, cement,
and glass industries. Characteristics of the three described mem-
brane applications as well as a few examples of ceria-based membrane
materials considered for each specific application are summarized in
Table 12.5.

12.6.4 Ceria membranes


In the previous section three different applications of oxygen mem-
branes were described and some of the material requirements
imposed by the operating conditions were outlined. Most of the
studies on mixed conducting oxygen membrane materials for the
above applications deal with single-phase perovskite or K2NiF4-type
materials such as (La, Sr)CoO3, (La, Sr)(Fe, Co)O3, (La, Sr)(Fe, Ga)O3,
(Ba, Sr)(Co, Fe)O3, and La2NiO4 and the reported performances
have been reviewed.393,396,397 Though excellent performance in terms
of flux have been demonstrated in many studies of single-phase per-
ovskite membranes,393,398,405–411 problems have also been reported
including kinetic demixing during operation,409 crack formation

b1469_Ch-12.indd 707 4/8/2013 12:41:02 PM


b1469 Catalysis by Ceria and Related Materials

708 C. Chatzichristodoulou et al.

Table 12.5 Characteristics of the three membrane applications shown in Fig. 12.20
and examples of ceria-based membrane materials considered for each specific
application.
Application Oxyfuel three end Oxyfuel four end Syngas
High pO2 5–20 bar 0.2–20 bar 0.2 bar
Low pO2 1 mbar–1 bar 30 mbar 10−19 bar
CO2, sulfur Not required Required Required
tolerance
Single-phase Ce1−yPryO2−x, Ce1−yPryO2−x, Ce1−yGdyO2−y/2
ceria system Ce1−yTbyO2−x, Ce1−yTbyO2−x, Ce1−ySmyO2−y/2
Ce1−y−zPryGdzO2−z/2−x Ce1−y−zPryGdzO2−z/2−x
Composite Ce1−yGdyO2−y/2– Ce1−yGdyO2−y/2– Ce1−yGdyO2−y/2–
systems (La, Sr)MnO3 (La, Sr)MnO3 (La, Sr)CrO3
Ce1−yGdyO2−y/2– Ce1−yGdyO2−y/2–
MnFe2O4 MnFe2O4
Ce1−yGdyO2−y/2–
(La, Sr)CoO3

due to chemical strain caused by loss of oxygen,397,412 phase transfor-


mations,413 and surface exchange limitations caused by carbonate
formation.414,415
Membranes based on acceptor-doped ceria may represent a
promising alternative to Ln,Tm-based perovskites. Ceria has a much
wider pO2 stability range and a higher ionic conductivity than most
of the Tm-based perovskites; at 700 °C the ionic conductivity of
Ce0.9Gd0.1O1.95 is ~ 0.04 S⋅cm−1,204,325 while the ionic conductivity of
La0.6Sr0.4CoO3, which is among the best ionic conductors of the
Tm-based perovskites, is ~ 0.016 S⋅cm−1.410 As discussed in Sections
12.4.2.1, 12.4.2.2, and 12.5.2.1, acceptor-doped ceria, over a very
wide pO2 and temperature range, is best described as an electrolyte
because its ionic conductivity by far exceeds its electronic conductiv-
ity; for example the pO2 range where ionic conductivity exceeds the
electronic is from pO2 ~ 10−20 atm to pO2 ~ 106 atm at 700 °C for
Ce0.9Gd0.1O1.95 (estimated from experimental data272). However,
n-type electronic conductivity is significant in acceptor-doped ceria
under reducing conditions (cf. Section 12.4.2.2 and 12.5.2.1).

b1469_Ch-12.indd 708 4/8/2013 12:41:02 PM


b1469 Catalysis by Ceria and Related Materials

Ceria and its Use in Solid Oxide Cells and Oxygen Membranes 709

Significant p-type electronic conductivity may be introduced at high


pO2 by doping with mixed-valence elements like Pr and Tb (cf.
Section 12.4.2.2), i.e. under specific conditions acceptor-doped
ceria is best characterized as a mixed conductor. Whereas nonnegli-
gible electronic conductivity is problematic when the material is
used as an electrolyte in SOFCs (as discussed in Section 12.5.2.1) it
opens the possibility of using the material for oxygen membranes.
In the following, different ways of achieving sufficient electronic
conductivity in acceptor-doped ceria for a membrane application
will be discussed. However, first the special case of the electrical
pumping of oxygen is considered.

12.6.4.1 Electrical pumping of oxygen


In the membrane applications discussed in the previous section and
summarized in Table 12.5, the driving force for oxygen permeation
is the oxygen activity difference between the two sides of the
membrane. Alternatively, where the oxide ion conductivity is much
higher than the electronic conductivity the oxygen flux can be
driven by applying an electrical potential difference over the
gastight layer. Typically, zirconia has been the material of choice for
such applications, but as the ionic conductivity of acceptor-doped
ceria is almost an order of magnitude higher than that of yttria-
stabilized zirconia at 600 °C, ceria has some advantages for this use.
Electrical pumping of oxygen is very energy consuming and is thus
primarily considered for specialized applications like the on-site
production of high-purity oxygen for welding, in research laborato-
ries, or military applications. Meixner et al.416 reports on results from
an Air Products/Ceramatec development project where a 32 cell
stack based on rare-earth-doped ceria was constructed and tested
over a period of 7000 h. The report is somewhat sparse on the
detailed conditions of operation and material compositions. The
cells were planar and were manufactured by casting. Interconnect
plates were made of Ca-doped lanthanum manganite (LCM) and
electrodes of (La, Sr)CoO3. It was demonstrated that the oxygen
flux was directly proportional to the applied current demonstrating

b1469_Ch-12.indd 709 4/8/2013 12:41:02 PM


b1469 Catalysis by Ceria and Related Materials

710 C. Chatzichristodoulou et al.

that electronic conductivity is small, and good gastightness was


achieved. The performance expressed as an area specific resistance
was quite stable over 7000 h. The ASR increased from ~ 0.56 Ω⋅cm2
to 0.62 Ω⋅cm2 over the test period corresponding to a degradation
rate of 1.6%/1000 h.416 Advantageously, the oxygen can be delivered
at a pressure exceeding atmospheric by enhancing the driving volt-
age. In a later paper from Ceramatec, Hutchings et al.417 demon-
strated the delivery of pure oxygen at a pressure of 2 MPa (20 atm)
for 5 cm2 planar cells based on thin (~ 50 µm) Ce0.845Sm0.15Co0.005O1.92
electrolytes that were co-fired with micro-channeled LCM “intercon-
nects”. La1−ySryCo0.2Fe0.8O3−x was used for the electrodes and a stable
operation for more than 3 years was documented when operating at
750 °C and 0.4 A⋅cm−2. Area specific resistances were of the order
~ 0.25 Ω⋅cm2 for small systems. (In passing it can be noted that for
tubular devices even higher oxygen delivery pressures of 17 MPa
have been demonstrated.418) These results show that durable com-
ponents based on thin-film ceria are indeed realizable.

12.6.4.2 High-temperature operation


As discussed in Section 12.4.2.2 of this chapter, classical low-
temperature electrolyte materials such as Ce0.9Gd0.1O1.95, Ce0.8Gd0.2O1.9,
and Ce0.9Sm0.1O1.95 acquire significant electronic conductivity when
at oxygen activities and temperatures where Ce(IV) starts reducing
to Ce(III). At 700 °C the pO2 where electronic conductivity is compa-
rable to the ionic is ~ 10−20 atm. However, at 1400 °C the two conduc-
tivities are comparable at ~ 10−4 atm.419 Hence, under a modest pO2
difference over the membrane, say from 0.21 to 10−4 atm, one should
be able to sustain significant fluxes over a ceria membrane. This was
shown to be the case by Park and Choi,419 who studied Ce0.8Gd0.2O1.9
with a specific view to deoxidizing steel melts, which requires high
chemical and thermal stability. The study was carried out with 2-mm-
thick samples and the oxygen permeation fluxes were measured as a
function of temperature and pO2. At 1600 °C permeation fluxes of
0.7 ml⋅min−1⋅cm−2 and 6 ml⋅min−1.cm−2 were found with a pO2 of 10−4
atm and 10−8 atm at the permeate side, respectively. The flux exceeds
that of zirconia-based systems, which has been researched more

b1469_Ch-12.indd 710 4/8/2013 12:41:02 PM


b1469 Catalysis by Ceria and Related Materials

Ceria and its Use in Solid Oxide Cells and Oxygen Membranes 711

heavily for this application, by one to two orders of magnitude. In a


later study420 the authors reported on the need to strengthen the
Ce0.8Gd0.2O1.9 to prevent mechanical failure due to expansion on
reduction. For the purpose of improving mechanical strength and
toughness samples with 2–10% Al2O3 addition were studied. The
addition of alumina led to the formation of a second phase ((Gd,Ce)
AlO3) observed in the grain boundaries. The second phase slightly
reduced the observed oxygen fluxes but the samples with the addi-
tion of 10% alumina stayed intact under permeation tests and con-
ductivity measurements at conditions where Ce0.8Gd0.2O1.9 was
observed to fail by cracking.

12.6.4.3 Thin-film acceptor-doped membranes


The pressure-driven permeation experiments on Ce1−yGdyO2−y/2 dis-
cussed above were carried out on millimeter-thick pellet samples and
the flux was clearly limited by bulk diffusion.420 To achieve high fluxes
at lower temperatures requires that the membrane thickness is
reduced preferably to where the flux is limited by surface processes.
A number of studies report results of membrane experiments based
on thin-film ceria-based membranes. Yin et al. prepared 10–20-µm-
thick tubular Ce0.8Gd0.2O1.9 membranes.421 A porous tube of CeO2 pre-
pared by extrusion was used as a support and the membrane layer was
applied by air spraying. Oxygen exchange kinetics at the exterior sur-
face was promoted by applying a layer of La0.2Sr0.8CoO3. Below 750 °C
the coated membranes showed higher fluxes than the uncoated ones
whereas at a higher temperature no beneficial effect was observed. At
900 °C fluxes of the order of 0.5 ml⋅min−1⋅cm−2 were measured with air
outside the tube and a He sweep flow inside (pO2 not stated).
From model studies,325,328 fluxes 10–20 times higher should be
expected at a high temperature when reducing the pO2 on the per-
meate side in the regime where significant n-type conductivity occurs
in the material. Oxygen partial pressures of 5 × 10−18 atm (33% O2,
66% CH4) are appropriate for membranes for syngas production
from methane or for oxidative coupling reactions. High oxygen per-
meation fluxes of the order of 10–20 ml⋅min−1⋅cm−2 measured at
900 °C have recently been reported by Chatzichristodoulou et al.343

b1469_Ch-12.indd 711 4/8/2013 12:41:02 PM


b1469 Catalysis by Ceria and Related Materials

712 C. Chatzichristodoulou et al.

These results were obtained on planar membranes with a thin


Ce0.9Gd0.1O1.95 layer (30 µm) on top of a porous Ni–YSZ support and
were manufactured by tape casting and lamination.422 To promote
surface exchange reactions on both sides, a Ni–YSZ cermet electrode
was applied on the low pO2 side and a La0.6Sr0.4Co0.2Fe0.8O3–
Ce0.9Gd0.1O1.95 composite electrode was applied on the high pO2 side.
The 5 × 5 cm2 test membranes were tested both with hydrogen/
steam mixtures and methane/steam mixtures fed to the permeate
side. At 900 °C a flux of 16 ml.min−1⋅cm−2 was observed, with air on
the high pO2 side and a 95%/5% hydrogen/steam mixture on the
low pO2 side. Comparing the experimental results with model pre-
dictions328,343 it was found that the flux is primarily limited by the
electronic conductivity of the membrane. Syngas production was
also demonstrated and post-test characterization of the membrane
showed no changes in membrane microstructure or other signs of
deterioration. These experiments clearly demonstrate the large
potential of ceria membranes for wide pO2 applications. Detailed
modeling estimated the oxygen stoichiometry profile in the mem-
brane during the test and specifically the vacancy concentration at
the low pO2 side. At the given operating conditions this was six times
less than if the material had been in equilibrium with the gas mix-
ture. The material is thus kinetically “protected” by the permeation
flux from the high pO2 side, which limits the expansion of the mate-
rial and thus the stresses in the membrane during use. As pointed
out in the previous section and discussed in several papers,16,420,423
stresses due to expansion related to vacancy formation in Ce1−y
GdyO2−y/2 is a major concern for ceria membranes. However, the
experiment discussed above328,343 illustrates that mechanically “safe”
operation is possible under application relevant conditions.

12.6.4.4 Enhancing electronic conductivity in ceria-based


membranes
As pointed out above, the performance-limiting factor of a Ce1−y
GdyO2−y/2 membrane is electronic conductivity. Different ways of
enhancing electronic conductivity in acceptor-doped ceria have

b1469_Ch-12.indd 712 4/8/2013 12:41:02 PM


b1469 Catalysis by Ceria and Related Materials

Ceria and its Use in Solid Oxide Cells and Oxygen Membranes 713

been pursued in order to obtain higher fluxes. These can roughly


be grouped into three categories, which are briefly presented below,
and are then discussed in turn in the three following sections:

• Composite membranes, where the membrane is a two-phase system


with one phase for ionic transport (ceria) and the second for
conduction of electrons. The latter could be a Tm-oxide such as
(Mn, Fe)3O4,424 a suitable perovskite,425,426 or a noble metal. Both
phases must be percolating.
• Aliovalent substitution, single-phase systems. The electronic conduc-
tivity of ceria can be increased by suitable substitution.
Substitution of Tb and Pr will lead to significant p-type electronic
conductivity at high pO2 (pO2 > 10−5 atm at 800 °C)98,264,266,269,272
and substitution with e.g. Zn has been reported to increase
n-type conductivity at low pO2 compared to Gd-doped ceria.427
• Grain structure modified acceptor-doped ceria, Co addition. In both pure
and acceptor-doped ceria, as discussed in Section 12.4.2.3, it is
well established that there is a significant enhancement in n-type
electronic conduction when the grain size is reduced to a nanom-
eter scale.141,279 Moreover, in several studies it has been demon-
strated that the addition of small amounts of Co-oxide270,271,428,429
increases electronic conductivity in the vicinity of the grain
boundaries.

Recorded membrane performances of a range of different mate-


rials systems ceria are summarized in Table 12.6. The materials have
been grouped into four classes: (i) composite membranes, (ii) mate-
rials with redox active dopants, (iii) acceptor-doped materials, and
(iv) materials with additional co-substitution. Selected results are
either discussed above (Sections 12.6.4.2 and 12.6.4.3) or in the fol-
lowing sections and in Section 12.6.4.5.

Composite membranes
Instead of requiring high mixed conductivity of the membrane
material one could achieve high permeability by combining a good

b1469_Ch-12.indd 713 4/8/2013 12:41:02 PM


b1469_Ch-12.indd 714

714
Table 12.6 Recorded membrane performances of a range of different materials involving ceria.

Flux (µmol. T
Materials Geometry L (µm) cm−2⋅s−1) (°C) Gas pO2′/pO2″ Ref.
Composite membranes

b1469
1 Ce0.8Gd0.2O1.9– Disk 600 0.08 950 air/He 430
La0.7Sr0.3MnO3 0.21/0.01
2 Ce0.8Gd0.2O1.9– Disk 1000 0.040 950 air/He 426

Catalysis by Ceria and Related Materials


C. Chatzichristodoulou et al.
La0.7Sr0.3MnO3
Ce0.8Gd0.2O1.9– Disk 1000 0.063 950 0.2/0.01 air/He
La0.8Sr0.2Fe0.8Co0.2O3 0.2/0.01
3 Ce0.8Gd0.1Co0.1O2 Disk 1250 0.002 a 850 air/He 426
0.2/0.01
4 Ce0.4Gd0.2Mn0.4O2−x Disk 1000 0.002 b 850 air/He 426
0.2/0.01
5 Ce0.75Nd0.25O1.875– Disk 600 0.2 900 0.1/0.003 431
Nd1.8Ce0.2CuO4 1030 0.12 900 0.1/0.003
1030 0.07 900 0.1/0.01
1030 0.04 850 0.1/0.01
1030 0.02 800 0.1/0.01
6 Ce0.8Sm0.2O1.9– Disk 300 0.14 0.21/0.01 432
La0.8Sr0.2CrO3
(Continued )
4/8/2013 12:41:02 PM
b1469_Ch-12.indd 715

Table 12.6 (Continued )


Flux (µmol. T
Materials Geometry L (µm) cm−2⋅s−1) (°C) Gas pO2′/pO2″ Ref.
7 Ce0.9Sm0.1O1.95 – Planar 300 1.1c 1000 air/He 424

Ceria and its Use in Solid Oxide Cells and Oxygen Membranes
Mn0.5CO1.5Ni0.5O4 Planar 300 7 air/Ar,CH4 air/
Ce0.9Sm0.1O1.95– 300 6 1000 Ar,CH4

b1469
MnFe2O4 140 10 air/Ar,CH4
8 Ce0.8Gd0.2O1.9– Disk 500 0.55d 950 0.21/0.005 425

Catalysis by Ceria and Related Materials


Gd0.2Sr0.8FeO3 500 0.25 850 0.21/0.005
1000 0.3 950 0.21/0.005
1000 0.14 850 0.21/0.005
500 3.41 950 air/syngas
9 Ce0.85Sm0.15O1.925– Disk 500 0.34e 950 0.21/0.005 433
Sm0.6Sr0.4FeO3 2.7 950 air/syngas
10 Ce0.8Sm0.2O1.9– Disk 600 0.46 950 0.21/0.005 434
LaBaCo2O5
11 Ce0.8Gd0.2O1.9– Disk ~1 0.006 700 0.21/0.0001 435
CoFe2O4
12 Ce0.9Sm0.2O1.9– Tube 1100 0.86f 950 air/CO 436
La0.8Sr0.2CrO3 0.21/10−15
13 Ce0.8Sm0.2O1.9– Hollow 100–200 0.3 950 air/CO2 437
La0.8Sr0.2MnO3 Fiber 0.2/0.01
14 CeO2–La0.2Sr0.8CoO3 Tube 10 0.007g 850 air/He 438
30 cm 0.21/0.001

715
4/8/2013 12:41:02 PM

(Continued )
b1469_Ch-12.indd 716

716
Table 12.6 (Continued )

Flux (µmol. T
Materials Geometry L (µm) cm−2⋅s−1) (°C) Gas pO2′/pO2″ Ref.
Redox active dopants

b1469
15 Ce0.6Zr0.2Pr0.2O2−x Disk 1000 0.005 700 air/He 266
0.015 950 0.21/10−4

Catalysis by Ceria and Related Materials


C. Chatzichristodoulou et al.
16 Ce0.8Pr0.2O2−x Disk 1000 0.008 800 0.21/0.021 267
0.012 850 0.21/0.021
0.019 900 0.21/0.021
0.025 950 0.21/0.021
0.05 1000 0.21/0.021
Purely acceptor-doped ceria
17 Ce0.8Gd0.2O1.9 2000 1 1400 0.21/10−8 419
Disk 0.1 0.21/0.015
18 Ce0.8Gd0.2O1.9 20 0.13 h 700 air/He 438
Tube 0.2 900 0.2/0.0001
19 Ce0.9Gd0.1O1.95 27 1.4 i 700 0.21/3.8 × 10−24 343
4.1 800 0.21/1.1 × 10−21
Planar 10.9 900 0.21/1.3 × 10−19

(Continued )
4/8/2013 12:41:02 PM
b1469_Ch-12.indd 717

Table 12.6 (Continued )

Flux (µmol. T
Materials Geometry L (µm) cm−2⋅s−1) (°C) Gas pO2′/pO2″ Ref.
Redox active dopants + Co

Ceria and its Use in Solid Oxide Cells and Oxygen Membranes
20 Ce0.8Pr0.2O2−x + 2% Co Disk 1000 0.02 j 700 0.21/0.021 429
0.04 800 0.21/0.021

b1469
0.06 850 0.21/0.021
0.08 950 0.21/0.021

Catalysis by Ceria and Related Materials


21 Ce0.8Tb0.2O2−x + 2% Co Disk 1200 0.054 1000 air/Ar 271
Ce0.8Tb0.2O2−x + 2% Co 1200 0.045 950 0.21/0.0001
Ce0.8Tb0.2O2−x + 2% Co 1200 0.035 900 0.21/0.0001
Ce0.9Tb0.1O2−x + 2% Co 900 0.013 1000 0.21/0.0001
Ce0.9Tb0.1O2−x + 2% Co 900 0.3 1000 air/CH4
Ce0.9Tb0.1O2−x + 2% Co 900 0.2 900 air/CH4
Notes: a The formula gives the nominal composition of the membrane material tested. 10 mol% Co is, however, beyond the solubility limit and
the membrane is in reality a two-phase mixture.
b
The formula gives the nominal composition of the membrane material tested. 40 mol% Mn is, however, beyond the solubility limit and the
membrane is in reality a two-phase mixture.
c
This study was carried out on a fairly large area cell (5 × 5 cm2), however, over 24 h only.
d
The authors report on a partial decomposition of the material after long-term testing.
e
No degradation was observed after 600 h of test.
f
Successful testing carried out for 1000 h.
g
The CeO2 is added here more to TEC match the composite membrane layer to that of the CeO2 support structure than to function as an ionic
conductor.
h
In this study porous electrode and catalyst layers were applied on both surfaces of the membrane to enhance surface exchange rates.

717
4/8/2013 12:41:02 PM

i
In this study porous electrodes were present on both membrane surfaces to enhance surface exchange kinetics.
j
Note that this value is an extrapolated estimate. The temperature interval in the experimental study is 700–850 °C.
b1469 Catalysis by Ceria and Related Materials

718 C. Chatzichristodoulou et al.

electronic conductor with a good oxide ion conducting electrolyte.


This was first suggested by Mazanec et al.439 in 1992 and in 2000
Kharton et al.430 reported membrane performances for various com-
posites based on acceptor-doped ceria as the oxide ion conducting
phase. Ceria was combined with classical electronic conducting
materials like (La, Sr)MnO3, (La, Sr)CoO3 and various spinels.426
Several studies exist440–442 where ceria-based composites have
been assessed for membrane use by measurements of total conduc-
tivity, ionic conductivity, and in some cases also the surface exchange
rate constants. It is well documented that improved total conductiv-
ity can be achieved by mixing ceria with a good electronic conductor
such as Co3O4440 or (La, Sr)CoO3442 and exchange kinetics can cer-
tainly be enhanced compared to the ceria only case,440,441 and in
some cases also beyond the values known for either of the phases
when studied as single phase.442
In several studies where a range of different mixing ratios
between the electronic and ionic conducting phase have been stud-
ied it has been found that an optimum composition exists.434,440,442
Typically, more of the ionic conducting phase is present at the opti-
mum as this is the poorer conductor of the two. Adding too much
of the electronically conducting phase may reduce the overall ionic
conductivity. The exact amount of the two phases for optimal per-
formance depends on microstructure, particle sizes as well as on
both electronic and ionic transport properties of both phases. In a
study of (Ce, Gd)O2 – (La,Sr)CoO3 composites, 20% (La, Sr)CoO3
(by volume) was needed to ensure significantly enhanced electronic
conductivity;440 it was found that 15% (by volume) of the electronic
conductor in a (Ce, Gd)O2–CO3O4 composite is enough for percola-
tion. With these compositions no negative effect was observed for
ionic conductivity.440,442 In a study with Sm-doped ceria 40%
LaBaCo2O5 (by volume) was needed to establish effective percola-
tion.434 In the following we shall limit the discussion to studies where
permeation experiments have been carried out. Such measure-
ments are more difficult to carry out than total conductivity meas-
urements, but they are more relevant from a performance assessment
point of view as they are much closer to the targeted application.

b1469_Ch-12.indd 718 4/8/2013 12:41:02 PM


b1469 Catalysis by Ceria and Related Materials

Ceria and its Use in Solid Oxide Cells and Oxygen Membranes 719

In an early work of Kharton et al. reporting on oxygen permea-


tion fluxes of composite membranes, (La,Sr)MnO3 was chosen as
the electronic conducting phase.430 The study used thick disks (L ~
0.5–1 mm) and demonstrated that indeed such materials could be
combined in an integral piece and sintered to gastightness. Fluxes
of the order 0.08 µmol⋅cm−2⋅s−1 at 950 °C were reported for 0.6-mm-
thick samples (log(pO2″/pO2′) ~ 1.2) and it was demonstrated that
the flux scales linearly with the driving force at high pressure (10–50 atm).
The fluxes, however, strongly decreased over time to ~ 10% of the
original level over 600 h,430 which is ascribed to detrimental reac-
tions between (La,Sr)MnO3 and ceria (e.g. interdiffusion of Mn in
ceria and the formation of SrCeO3 at grain boundaries). Later, a
broader range of electronically conducting materials was investi-
gated as the electronic conducting second phase in the compos-
ite.426 It was demonstrated that slightly superior permeation
properties could be achieved using La0.8Sr0.2Fe0.8Co0.2O3 and more
importantly that the flux was stable over time (demonstrated over
200 h). Interestingly, the best permeation properties of the compos-
ites investigated in Kharton et al.426 were for a membrane made of
Ce0.8Gd0.1Co0.1O2. The added 10% Co is above the solubility limit in
ceria and below the normal percolation threshold if two phases of
particulate origin are randomly mixed, clearly indicating a benefi-
cial effect of adding small amounts of Co to ceria for membrane
purpose, as will be discussed further in the section below.
Since the early works of Kharton et al., a significant number of
different composite systems based on ceria have been investigated.
Measured fluxes for a selected subset are reproduced in Table 12.6,
which gives measurements obtained both on disk-shaped as well as
tubular or hollow fiber-shaped samples with various thicknesses. The
table shows flux values obtained over a narrow pO2 span typically
experimentally arranged with air on one side of the membrane and
a He purge flow on the other, establishing a pressure ratio of the
order of 0.21/0.005 to 0.21/0.02, as well as fluxes measured over a
wide pO2 span with air on one side and CO, H2/H2O, or syngas on
the low pO2 side defining a pressure ratio of 0.21/10−15 to 0.21/10−20.
A detailed comparison of the flux values obtained for the various

b1469_Ch-12.indd 719 4/8/2013 12:41:02 PM


b1469 Catalysis by Ceria and Related Materials

720 C. Chatzichristodoulou et al.

systems would require a detailed correction for differences in tem-


perature, membrane thickness as well as pressure gradient over the
membrane between the different experiments, which is beyond the
scope of this work. The overall picture from the values reported in
Table 12.6 is that the fluxes measured over a wide pO2 span are typi-
cally 6–8 times larger than over a narrow span at the same tempera-
ture, which roughly corresponds to the expectation considering the
change in driving force. Also it seems that flux levels within the
investigated thickness range scale roughly in inversely proportion to
the thickness. Both these observations indicate that flux values are
primarily controlled by bulk transport (cf. Eq. (12.42)).
The flux values summarized in Table 12.6, despite the difficul-
ties in achieving a fair comparison between the different systems,
indicate that the systems: Ce 0.85Sm0.15O1.925–Sm0.6Sr0.4FeO3,
Ce0.9Sm0.1O1.95–MnFe2O4 (85–15), and Ce0.8Sm0.2O1.9–LaBaCo2O5
(60–40) are the more promising ones, which for further work high-
lights, that: (i) good performance can indeed be achieved with spi-
nel-type second phases,424 (ii) the strategy of letting the acceptor
dopant of the ceria be the A-site element in the perovskite second
phase, as in Zhu et al.,425,433 seems fruitful for mitigating the detri-
mental effects of reactions between the phases, and finally (iii) that
some ionic conductivity in the primarily electronically conducting
part of the composite is beneficial as seen for the Ce0.8Sm0.2O1.9–
LaBaCo2O5 system434 and also pointed out in studies of (La,Sr)
CoO3–Ce0.9Gd0.1O1.95442 and (La,Sr)(Co,Fe)O3–Ce0.8Gd0.2O1.9.426
From the work on composite membranes it is clear that the con-
cept has some advantages; sufficient electronic conductivity can
indeed be introduced by the addition of a second phase, which fur-
ther allows the thermal expansion coefficient (TEC) of the mem-
brane to be tailored. Moreover, the addition of a second phase or
elemental substitution in ceria may in some cases improve robust-
ness by mitigating the mechanical consequences of the expansion
on reduction of the ceria phase.443,444 However, the concept also
presents some difficulties, notably being the suppression of reac-
tions between the two phases both during manufacture and use as
well as ensuring gastightness and mechanical integrity. Whereas

b1469_Ch-12.indd 720 4/8/2013 12:41:02 PM


b1469 Catalysis by Ceria and Related Materials

Ceria and its Use in Solid Oxide Cells and Oxygen Membranes 721

maximizing permeation by using supported thin-film membranes


has been demonstrated for single-phase membranes this strategy for
performance optimization has not yet, to our knowledge, been dem-
onstrated for a two-phase composite. Most studies of composite
membranes have been carried out on fairly small (~ 1 cm diameter)
flat disk-type samples with thicknesses in the range from 300 to
1000 µm. In a few cases some upscale in the direction of what would
be required for technological use of such systems has been
attempted. Takamura et al.444 showed that it is possible to make
25 cm2 area pieces with a membrane thickness of only 133 µm and
tested a module with an active area of 13 cm2 involving metal mani-
fold plates.444 However, long-term durability data were not reported
for the system. Wang et al. reported on testing periods of over 1000 h
demonstrating the structural stability of a small tubular composite
membrane under appropriate conditions for syngas production436
and Zhu et al.433 demonstrated stable fluxes under syngas conditions
for a period of ~ 500 h. Both these studies are among those showing
good membrane performance in terms of flux and they clearly show
the potential of the concept. More work is needed to demonstrate
both improved performance, upscale in terms membrane area, and
long-term durability.

Single-phase membranes: substitution with redox active elements


Under certain conditions acceptor-doped ceria itself exhibits signifi-
cant electronic conductivity and addition of an electronic conducting
second phase is not necessary. This is for instance the case with
the classical acceptor-doped materials Ce1−yGdyO2−y/2 and Ce1−y
SmyO2−y/2 under wide pO2 membrane applications (with strong
reduction on the low pO2 side). Also, other redox active elements
may be substituted in ceria to enhance further either the p-type or
n-type conductivity. The former can be achieved with e.g. the addition
of Pr or Tb, and numerous studies exist for Pr substitution258,259,263,269,272
or Pr and Gd substitution261,266,445 (see also Section 12.4.2.2). Fewer
studies are available where enhanced n-type conduction has been
introduced by substitution with redox active elements, but recently

b1469_Ch-12.indd 721 4/8/2013 12:41:02 PM


b1469 Catalysis by Ceria and Related Materials

722 C. Chatzichristodoulou et al.

Schmale et al.270 reported enhanced n-type conductivity by partial


substitution with Zn.

Pr/Tb substituted ceria; conductivity studies


Numerous studies exist on the fundamental transport properties
(ionic and electronic conductivity) of Pr/Tb substituted ceria,
which show that such materials could be used for oxygen mem-
branes. Takasu258 and later Nauer259 investigated the conductivity of
Pr-doped ceria and reported a significant electronic contribution to
the total conductivity. Shuk and Greenblatt263 investigated a range of
different Pr substitutions with respect to ionic and electronic con-
ductivity and were among the first to point to the potential applica-
tion of suitably doped ceria in an oxygen membrane, which was
elucidated by calculating expected fluxes as deduced from the meas-
ured conductivities (electronic and ionic). The flux predicted for a
100-µm membrane of Ce0.7Pr0.3O2−x at 700 °C is ~ 0.02 µmol⋅cm−2⋅s−1
between oxygen and air. The fundamental charge transport proper-
ties of Pr-doped ceria have already been discussed in Section 12.4.2.2.
One potential problem with the Pr-substituted materials is the large
apparent TEC values, which may complicate their practical techno-
logical use as they are prone to mechanical failure. For example the
thermal expansion coefficient of Ce0.8Pr0.20O2−x varies gradually
between 11–14 × 10−6 K−1 from room temperature to 400 °C, where it
increases dramatically to reach a maximum of ~ 40 × 10−6 K−1 at
675 °C after which it decreases again to around 20 × 10−6 K−1 at
1000 °C. This “anomaly” is associated with the valence change of Pr17
and is thus an inherent problem for fluorites with redox active
dopants. Due to this, it would be of great interest to study the effect
of small additions of redox active elements to acceptor-doped ceria.
The effects of Pr when added to ceria in combination with other
elements (other acceptor dopants) have been studied for several
different materials. Lübke and Wiemhöfer261 showed that co-doping
3% Pr into ceria with 17% Gd increased the p-type conductivity at
high pO2 by a factor of ~ 10 at 750 °C and suppressed the n-type con-
ductivity at low pO2. In good agreement, Kharton et al.264 found that

b1469_Ch-12.indd 722 4/8/2013 12:41:02 PM


b1469 Catalysis by Ceria and Related Materials

Ceria and its Use in Solid Oxide Cells and Oxygen Membranes 723

p-type conductivity for 2% Pr and 18% Gd substituted ceria was


enhanced by a factor of 2.5–4 compared to the 20% Gd substituted
material. Further, they showed that the Pr substitution did not
penalize ionic conductivity.

Effect of adding Co
Numerous studies exist on the possibility of enhancing electronic
conductivity even further in these compounds by the addition of
Co.180,270,428 Co is known to be a very effective sintering aid to
ceria180,428 capable of reducing the sintering temperatures from
around 1500 °C to around 1000 °C. The solubility limit of Co in ceria
may be as low as 0.5%;446 however, usually more Co than this is used
(~ 2%).270,271,429 It is by now fairly well established that the excess Co
forms an electronically conducting phase along the grain boundaries
enhancing the p-type electronic conductivity in the samples.270,429,446
In a study of Ce1−yPryO2−x (CPO) (20% Pr) with the addition of
0, 2, or 5% Co, Fagg268 found that the average p-type conductivity
was enhanced by the addition of 2% Co from ca. 0.018 S.cm−1 to
0.061 S⋅cm−1 at 950 °C when assessed between 1 to 0.2 atm of oxygen
partial pressure (EMF method). At 700 °C the addition of Co
increased the electronic conductivity from 0.02 to 0.04 S⋅cm−1. An
addition of 2% does not penalize the ionic conductivity whereas the
addition of 5% Co results in reduced ionic conductivity without
enhancing the electronic conductivity any further. Schmale et al.270
studied the effect of adding 2% Co to the electronic transport prop-
erties of Gd/Pr-doped ceria. The p-type conductivity increased with
Pr content over the series Ce0.8Gd0,2−yPryO2−x (y = 0.05, 0.1, 0.15). For
the Gd-rich sample a significant increase in electronic conductivity
was observed over the entire pO2 range studied from log pO2 = −0.6
to log pO2 = −13 at 700 °C, 750 °C, and 800 °C, i.e. an increase both
in the n-type region and the p-type region. For the Pr-rich sample,
Co increased the electronic conductivity over the entire pO2 range
only at 800 °C, whereas at 750 °C and 700 °C the electronic conduc-
tivity was enhanced at low pO2 (log pO2 < −7.5) but reduced at high

b1469_Ch-12.indd 723 4/8/2013 12:41:02 PM


b1469 Catalysis by Ceria and Related Materials

724 C. Chatzichristodoulou et al.

pO2. The authors ascribe the additional conductivity contribution to


a “percolating” Co-rich phase along the grain boundaries.
Balaguer271 also found enhanced electronic conductivity with
the addition of 2% Co to Tb-substituted ceria (Ce1−yTbyO2−x y = 0.1
and 0.1). At 800 °C in oxygen the electronic conductivity was
enhanced by a factor of 1.5–2.5.

12.6.4.5 Membrane experiments with redox active dopants


As expected from the discussion in the previous section, the elec-
tronic conductivity in acceptor-doped ceria (e.g. Sm-doped or
Gd-doped ceria) and ceria doped with redox active elements such as
Pr or Tb, as well as for materials with multiple substituents (e.g. Gd
and Pr substitution) may be sufficient for use as an oxygen mem-
brane. Several studies give measurements of the oxygen fluxes
through dense samples of doped ceria. A subset of measured oxygen
fluxes as well as the experimental conditions under which they were
measured is collected in Table 12.6. The performance of the purely
acceptor-doped materials (materials 17, 18, and 19) was discussed in
Section 12.6.4.3. The performance of membrane materials with
redox active dopants (materials 15 and 16) and with redox active
dopants + Co addition (materials 20 and 21) is discussed in this
section.
In an early study by Qi et al.266 achievable fluxes through dense
single-phase ceria membranes were studied. Several different mate-
rials with different amounts of Pr, Sm, Tb, and Zr substitutions as
well as combinations were studied in the form of thick membrane
disks at fairly high pO2 (air on the high pO2 side and a He purge on
the low pO2 side). From the measured conductivities Ce0.6Zr0.2Pr0.2O2−x
was considered the most promising material and was tested for oxy-
gen permeation. At 950 °C, fluxes of the order of 0.015 µmol⋅cm−2⋅s−1
were measured (see Table 12.6, material 15) this is a factor of 4–20
lower than values measured for different composite ceria-based
membranes under similar conditions (Table 12.6, materials 2, 8, 9,
and 10). Fagg et al.267 also measured permeation fluxes for Pr and
Pr/Zr substituted ceria. The best performance in terms of flux was

b1469_Ch-12.indd 724 4/8/2013 12:41:02 PM


b1469 Catalysis by Ceria and Related Materials

Ceria and its Use in Solid Oxide Cells and Oxygen Membranes 725

found for the purely Pr-substituted materials, where fluxes of the


order of 0.025 µmol⋅cm−2⋅s−1 were measured at 950 °C on 1-mm disks
for log pO2,s″/pO2,s′ ~ 1. The addition of Zr was found to reduce the
fluxes but also the material expansion related to loss of oxygen,
making the material less prone to mechanical failure.
As discussed in the previous section, the addition of small
amounts of Co to acceptor-doped ceria increases the electronic con-
ductivity and hence should also be expected to increase the achiev-
able oxygen permeation flux in a membrane experiment. This has
been demonstrated to be the case for both Pr-doped ceria429 and
Tb-doped ceria.271 As expected, the effect is most pronounced at low
temperatures. Fagg showed that fluxes increase by a factor of ~ 5
at 850 °C from ca. 0.012 µmol⋅cm−2⋅s−1 to 0.06 µmol⋅cm−2⋅s−1
(cf. Table 12.6 samples 16 and 21) with the introduction of Co.
Similarly, Balaguer271 reported fairly high fluxes ~ 0.035 µmol⋅cm−2⋅s−1
at 900 °C and 0.05 µmol⋅cm−2⋅s−1 at 950 °C for thick pellet samples
(1 mm) of Tb-doped ceria with the addition of 2% Co. A character-
istic of membranes with enhanced electronic conductivity via Co
addition is a weaker temperature dependence with an activation
energy around 50 kJ⋅mol−1,429 compared to ~ 100 kJ⋅mol−1 for purely
Pr-doped samples. Co addition is thus effective for enhancing the
flux especially at low temperatures where flux levels for Pr-doped
materials with Co are at 850 °C comparable to flux levels measured
in (Ce, Gd)O2–(La, Sr)MnO3 composites and about a factor of two
lower than for the best performing composite systems.425

12.6.5 Mechanical problems of ceria in membranes


When ceria is used as a membrane, significant stoichiometry varia-
tions may be induced over the membrane.422 As the unit cell volume
of ceria depends on the oxygen content (cf. Section 12.2.2.3) this
results in a strain profile across the membrane, which leads to
mechanical stresses. The strain associated with stoichiometry
changes may, as discussed in Section 12.2.2.3, be of the order of
1–2%. Typical rupture strain levels in ceramics are of the order of
0.1%. This means that these lattice strains can result in mechanical

b1469_Ch-12.indd 725 4/8/2013 12:41:02 PM


b1469 Catalysis by Ceria and Related Materials

726 C. Chatzichristodoulou et al.

failure, if for example (i) a significant stoichiometry gradient exists


over the material, (ii) the material is constrained on a support that
does not change in size with pO2, or (iii) the sample is clamped at
the edges. The mechanics of this situation have been analyzed by
Atkinson423 and Hendriksen et al.412 The stress levels are dependent
on the actual stoichiometry variation realized over the membrane
during use, which, as also discussed in the previous section, depends
not only the conditions of operation (T, pO2′, pO2″) but also on the
electrode polarization resistances. For the thin-film membrane
experiment reported in Chatzichristodoulou et al.343 and Kaiser
et al.422 it was assessed that the expansion encountered under mem-
brane operation is six times less than if the material had been in
equilibrium with the gas phase on the low pO2 side, which limited
the strains to a tolerable level.
Besides limiting the stress level by keeping the material suffi-
ciently oxidized during operation as discussed above, the risk of
failure can be minimized by proper design of the membrane mod-
ule: the module should be designed such that the membrane layer
is always in compression and the risk of a buckling-driven delamina-
tion should be minimized by reducing the membrane thickness.328,412
Finally, the stress levels can to a minor degree be minimized via the
choice of dopant, i.e. Tb/Pr-co-doped materials expand less for a
given change in oxygen stoichiometry than does a singly Pr- or
Tb-substituted material.17

12.6.6 Concluding remarks on ceria membrane performance


The performance of acceptor-doped ceria membranes, composite
ceria-based membranes, and membranes with redox active dopants
was reviewed in Sections 12.6.4.2, 12.6.4.3, 12.6.4.4, and 12.6.4.5,
respectively. Variations in sample geometry, membrane thickness,
as well as test conditions make comparison of performance some-
what difficult. Also for some studies, catalysts have been added to
the surface of the samples to enhance the flux, whereas in other
studies the as-prepared surface is used in the experiments.
However, in particular, studies of thin-film membranes, both the

b1469_Ch-12.indd 726 4/8/2013 12:41:02 PM


b1469 Catalysis by Ceria and Related Materials

Ceria and its Use in Solid Oxide Cells and Oxygen Membranes 727

supported 30-µm Ce0.9Gd0.1O1.95 membranes343 and the 140-µm


composite (Ce,Sm)O2–MnFe2O4 membrane,447 demonstrate that
very significant permeation rates can be achieved with ceria-based
membranes both considering narrow pO2 range applications (oxy-
fuel) as well as large gradient applications (syngas), cf. Table 12.6.
It is clear that to achieve substantial permeation rates in ceria-based
systems the following conditions should be fulfilled: (i) the mem-
brane layer should be thin, (ii) good catalysts must be applied to
both surfaces, and (iii) the electronic conductivity must be ensured
either by (a) strong reduction from the low pO2 side, (b) by the
addition of a second phase to carry the electrons, or (c) suitable
doping with redox active species (Gd+Pr/Tb + Co-addition). The
enhancement of electronic conductivity by Pr or Tb substitution
and Co addition has been clearly demonstrated in several studies
but membrane measurements271,429 on these systems have so far
been limited to thick-film samples resulting in flux levels not
exceeding 0.5 µmol cm−2⋅s−1.
A complicating factor for the use of ceria-based membranes is
the large chemical expansion that occurs on oxygen loss from the
material, which may result in mechanical failure (cf. Section 12.6.5)
under specific conditions of use. The successful operation417 of a
ceria-based membrane for the electrochemical pumping of oxygen
over several years demonstrated that the associated stresses under
small pO2 gradients (oxygen/air) certainly need not be detrimental
and successful membrane operation under large pO2 gradients
(air/syngas) has also been documented recently.343
In summary, ceria has a great potential for use as an oxygen
membrane. Significant fluxes can be obtained when thin membranes
are manufactured and suitable electrodes applied. For wide pO2
applications (syngas and the oxidative coupling reaction) acceptor-
doped (Ce,Gd)O2, (Ce,Sm)O2 can be applied and flux levels are
comparable to levels reported for thin-film perovskite membranes.393
Performance may be improved if additional n-type conductivity can
be introduced by doping. For applications under oxidizing condi-
tions (oxyfuel) electronic conductivity must be enhanced by addition
of a second phase or suitable doping. Also, under such conditions

b1469_Ch-12.indd 727 4/8/2013 12:41:03 PM


b1469 Catalysis by Ceria and Related Materials

728 C. Chatzichristodoulou et al.

(10−5 < pO2 < 0.2 atm) significant fluxes have been measured on com-
posite systems. Doping with Pr or Pr/Gd and addition of Co also
seems a very promising route, but so far nobody has reported flux
measurements on thin-film systems with these materials. The advan-
tage over the alternative perovskite-based membranes is the high
chemical stability of ceria, which will not form carbonates when
exposed to CO2 and is likely to show better tolerance to pollutants.

12.7 Ceria in Solid Oxide Fuel Cell Electrodes


12.7.1 Introduction
12.7.1.1 Solid oxide fuel cells and electrocatalysis
A solid oxide fuel cell (SOFC) is a high-temperature electrochemical
device (typically operated above 600 °C), which converts chemical
energy directly into electrical energy. An introduction to SOFCs was
given in Section 12.5.1. One of the advantages of SOFCs (compared to
low-temperature fuel cells) is the tolerance of nonhydrogen fuels, e.g.
carbon monoxide, which acts as a poison in low-temperature fuel cells.
In principle, all chemicals that can be combusted (oxidized) are poten-
tial fuels for SOFCs. The operating temperature of SOFCs is normally
sufficiently high to sustain the process of reforming hydrocarbons, e.g.
CH4, which is the fuel of preference for most applications. However,
the required steam reforming causes several problems in practice,
which has led to dedicated research on directly feeding the SOFC with
hydrocarbons without any pretreatment of the fuel (see Mogensen and
Kammer448 and references therein). Thus, there are a few alternative
anode reactions one could consider for SOFC operation:

Oxidation of carbon monoxide: CO + O2− → CO2 + 2e− (12.45)

Oxidation of hydrocarbons:
Cx H y + (2x - 2 )O2- Æ x CO2 + 2 H2O + (4x + y)e-
y y
(12.46)

As will be discussed later, a reaction like (12.46) is highly unlikely to


occur in one step even for the simplest hydrocarbon, CH4.

b1469_Ch-12.indd 728 4/8/2013 12:41:03 PM


b1469 Catalysis by Ceria and Related Materials

Ceria and its Use in Solid Oxide Cells and Oxygen Membranes 729

In all energy systems there are inherent losses in the energy con-
version process. This is naturally also valid for SOFCs where energy
is required to overcome kinetic barriers especially at the electrodes.
These losses are usually termed the overpotential, which is by defini-
tion the difference between the actual electrode potential at a given
current and the equilibrium (zero current) electrode potential, i.e.
the extra potential that must be applied to cause an electrode reac-
tion at a certain rate. It should be noted that the SOFC literature
often refers to the electrode polarization resistance instead of speci-
fying the reaction rate, because for SOFCs there is an approximately
linear relation between current density (reaction rate) and overpo-
tential. It is important during the reading of this section to bear in
mind that a low polarization resistance is equivalent to a high reac-
tion rate.
An electrode reaction may be anodic (oxidation) or cathodic
(reduction) depending on the sign of the overpotential with respect
to the equilibrium potential of the electrode. In Fig. 12.15, each of
the two electrodes and their associated half-cell reactions requires
their own specialized electrocatalyst. The simplest definition of an
electrocatalyst might be that it is a catalyst that participates in elec-
trochemical reactions. However, some catalysts that operate through
electrochemical reactions are not regarded as electrocatalysts.
These are the so-called Mars–van Krevelen type of catalysts449,450
where both anodic and cathodic reactions take place in parallel on
the surfaces of the catalysts. This means that these catalysts need not
be part of the electrochemical cell. The electrical potential of the
catalyst in this case is a mixed potential, the value of which is deter-
mined by the reaction conditions, and it is in general not known or
measured. Thus, a definition of an electrocatalyst as a catalyst that
participates in electrochemical reactions in electrodes of electro-
chemical cells seems more in line with the actual use of the term. It
should be further stressed that in electrocatalysis the equilibrium of
a reaction can be shifted by varying the potential of the electrode,
which is not possible in heterogeneous catalysis.
Great care should therefore be exercised when data from the
catalytic literature relating to the Mars–van Krevelen type of catalysts

b1469_Ch-12.indd 729 4/8/2013 12:41:03 PM


b1469 Catalysis by Ceria and Related Materials

730 C. Chatzichristodoulou et al.

are applied to electrocatalysis because the potential at which a


Mars–van Krevelen type of catalyst operates will most often be hun-
dreds of millivolts away from that of an electrocatalyst.
An electrocatalyst lowers the activation energy for an electro-
chemical reaction without altering the reaction equilibrium, similar
to a traditional catalyst. In practice, this is seen as a decrease in the
overpotential of the electrochemical reactions. The electrocatalyst
may constitute the entire electrode or be an addition to a composite
electrode. It need not necessarily be an electronic conductor, but
the electrode as such must naturally be electronically conducting.

12.7.1.2 Electrode requirements of a SOFC


• Catalytic activity The electrode must have a high (electro)cata-
lytic activity for the electrochemical oxidation of the fuel gas
(anode) or the reduction of oxygen (cathode).
• Impurity tolerance The electrodes should be tolerant to some level
of contaminants, e.g. sulfur and silicon, which are commonly
present in the supplied gasses or as impurities in the raw materi-
als used. The impurities can react with the component materials
or condense at the electrochemically active sites, affecting their
properties and eventually leading to reduced performance. If
hydrocarbon-based fuels are used, the anode should also be
resistant to coke formation.
• Stability The electrode must be chemically, morphologically, and
dimensionally stable in the gaseous environment and at the high
operating temperatures during long-term operation. On the
fuel side, the composition of the gas normally changes over the
cell, i.e. more fuel is consumed as the gas flows along the cell
producing more and more steam or CO2. The anode must there-
fore be stable, not only at the fuel inlet conditions but also at the
more oxidizing environment at the fuel outlet. There is a vary-
ing pO2 on the cathode side as well. Therefore, the cathode
should also be stable over a range of pO2 and overpotential.
• Conductivity High electrical conductivity, both electronic and
ionic, under a broad range of operating conditions is desired.

b1469_Ch-12.indd 730 4/8/2013 12:41:03 PM


b1469 Catalysis by Ceria and Related Materials

Ceria and its Use in Solid Oxide Cells and Oxygen Membranes 731

This is needed in order to facilitate the electrochemical reac-


tions and to collect/deliver the electrons/ions from/to the reac-
tion sites with minimal ohmic losses.
• Compatibility The electrode must be chemically, thermally, and
mechanically compatible with the other cell components, not
only at the operating conditions, but also during fabrication
where normally much higher temperatures and different gas
environments are used.
• Porosity The porosity of the electrode must be sufficient to allow
gas transport to and from the reaction sites and tailored with
regard to both mass transport conditions as well as adequate
mechanical strength.

Although a wide range of materials has been considered as elec-


trode materials for SOFCs, the state-of-the-art material used as a fuel
electrode in SOFCs (e.g. for electrochemical oxidation of hydrogen)
is a cermet of nickel metal in combination with yttria-stabilized zirco-
nia, Y2O3-doped ZrO2 (YSZ).1,293 Ni is used due to its excellent ability
to catalytically dissociate the most commonly used fuel in SOFCs, i.e.
hydrogen. It is also one of the metals that are able to withstand the
operating conditions of an SOFC, i.e. high temperatures (up to
1000 °C) and reducing atmospheres. Nickel, as a metal, also acts as an
electronically conducting phase that is needed to transport the elec-
trons from the reaction sites to the current collector. The function of
YSZ in the anode is to support the Ni particles, inhibit coarsening of
the metallic particles at the operating temperature, and to provide a
thermal expansion coefficient similar to other cell components. The
YSZ phase also forms conducting paths for oxide ion transport, and
thereby enlarges the electrode volume that is active for electrochemi-
cal reactions. Each of the three phases of an Ni–YSZ cermet, i.e. Ni,
YSZ, and the gas phase (porosity), should form percolating networks.
The electrocatalytically active area is a zone along the boundary
where the three phases meet called the triple-phase boundary (TPB),
see Fig. 12.21.
For the air electrode a composite comprising Sr-doped lantha-
num manganite (LSM) and YSZ has been the initial choice

b1469_Ch-12.indd 731 4/8/2013 12:41:03 PM


b1469 Catalysis by Ceria and Related Materials

732 C. Chatzichristodoulou et al.

Figure 12.21 (left) The triple-phase boundaries (circled areas) in an SOFC with
composite electrodes. (right) If using mixed ionic and electronic conducting
(MIEC) electrodes the reactions can in principle be spread out over a larger sur-
face as indicated by the outlined particles.

for commercial use. LSM has been used due to its compatibility, or
low-level reactivity, with YSZ electrolytes and due to its similar coef-
ficient of thermal expansion to YSZ. In spite of the differences in
material properties, the two types of electrode have some common
features, one being the composite nature of the electrodes with the
electrochemical reactions taking place near the TPB. LSM is an
oxide with a perovskite structure and it has high electronic conduc-
tivity, which confines the reduction of oxygen to O2− close to the
TPB. LSM is therefore analogous to Ni in the anode, being the
electrocatalyst (for the O2 reduction) and the electronic
conductor.451,452
Despite the relative success of using Ni–YSZ cermets as the
anode and LSM–YSZ composites as the cathode in SOFCs, these
electrodes suffer from several drawbacks. Below is a selection of
some of these shortcomings:

• Sulfur poisoning (both electrodes) Impurities in the fuel stream,


especially sulfur, inhibit anode performance. Strong reversible
poisoning of the Ni–YSZ anode occurs at feed concentrations of
1 ppm H2S in H2 at 1000 °C and as low as 50 ppb H2S in H2 at

b1469_Ch-12.indd 732 4/8/2013 12:41:03 PM


b1469 Catalysis by Ceria and Related Materials

Ceria and its Use in Solid Oxide Cells and Oxygen Membranes 733

750 °C.453 Even at these low sulfur concentrations, desulfuriza-


tion of the fuel gas is needed. Pretreatment of the fuel can solve
this problem, but a sulfur-tolerant anode would eliminate the
need for pretreatment, helping to reduce SOFC system costs. It
has recently been shown that trace amounts of SOX in air could
potentially also be a problem for cathodes during long-term
operation.454
• Oxidation-reduction intolerance (anode) Nickel oxidizes rapidly, and
expands by 69 vol% under oxidation.455 Because of the large
volume change during reduction/oxidation, the structure of
the anode and its mechanical strength are severely affected and
the Ni–YSZ anode heavily degrades during redox cycles.456 The
anode may break apart when exposed to air, which can
occur during start-up and shutdown of the fuel cell system. With
a Ni–YSZ supported cell the electrolyte will crack in the first
redox cycle. An anode that does not need protection from air
would simplify the system and lower costs.
• Thermal expansion mismatch (both electrodes) The TEC of both the
Ni–YSZ anode and the LSM–YSZ cathode is higher than that
of the electrolyte (YSZ-based) and other cell components. The
TEC varies with the composition of each electrode, increasing
with increased Ni content in the anode (see e.g. Minh and
Takahashi1 and references therein) or with increasing Sr-doping
in LSM.452,457 This can lead to mechanical stability problems
during e.g. thermal cycling of the fuel cell. However, on the
cathode side, the alternative cathode materials being developed,
such as cobaltites and ferro-cobaltites, have an even higher ther-
mal expansion mismatch.452,457
• Carbon deposition (anode) Another problem is the use of hydrocar-
bon-based fuels. It is widely agreed that direct feeding of dry
hydrocarbons into a fuel cell must be avoided when using a
Ni-based anode. Ni is an efficient catalyst for hydrocarbon crack-
ing, resulting in carbon deposition on the electrode.458 Coking
destroys the Ni–YSZ cermet mechanically (see e.g. Gorte and
Vohs459 and references therein). A significant amount of steam
must be added to induce steam reforming in order to prevent

b1469_Ch-12.indd 733 4/8/2013 12:41:03 PM


b1469 Catalysis by Ceria and Related Materials

734 C. Chatzichristodoulou et al.

coke formation at the Ni–YSZ anode. However, there are several


disadvantages to using reforming448 compared to using an elec-
trode that can tolerate dry hydrocarbons. The high steam con-
tent results in a diluted fuel, which has a negative effect on
efficiency. Additional equipment such as heat exchangers, steam
raising equipment, and exhaust gas recycling equipment adds
extra cost to an SOFC system.
• Chemical compatibility (cathode) Manganite-based cathode materi-
als have reasonable chemical compatibility with YSZ-based elec-
trolytes. However, high sintering temperatures (> 1200 °C) can
lead to the formation of electronically insulating SrZrO3 and
La2Zr2O7 phases.452,460–462 Introducing a slight (La, Sr) deficiency
in the materials can decrease the unwanted reactivity to some
extent.

The scientific community in collaboration with industry is constantly


trying to solve the problems mentioned above, which are inherent
to the materials used as typical components in SOFCs. The question
is: what can be done from a materials or electrode design point of
view to solve or circumvent these issues?
One advantageous design solution would be to extend the
length and width of the TPB zones as much as possible. One way of
achieving this is by optimizing the microstructure of the electrode
composite to increase the TPB length. However, the ideal electrode
would be a one-phase material with high mixed conductivity, i.e. an
electrode material that conducts both electrons and oxide ions. In a
MIEC, the TPB would not be limited to locations where the differ-
ent phases connect, i.e. the ionic conducting phase, the electronic
conducting phase, and the gas phase (see Fig. 12.21). Taking the
fuel electrode as an example, a MIEC material would provide both
oxide ions to the reaction sites and transport the released electrons
from the oxidation reaction. The fuel oxidation (at the anode) or
the oxygen reduction (at the cathode) could thus theoretically take
place over the entire surface of a MIEC electrode. The ratio between
the ionic conductivity (σi) and the electronic conductivity (σe) of a
MIEC is also of importance; it will influence where in the electrode

b1469_Ch-12.indd 734 4/8/2013 12:41:03 PM


b1469 Catalysis by Ceria and Related Materials

Ceria and its Use in Solid Oxide Cells and Oxygen Membranes 735

the actual reactions take place.463 However, it has proven very diffi-
cult to find a single material that fulfils all the requirements of a fuel
electrode.464 For the air electrode the situation is quite different.
There are a number of materials that have been explored in the
literature for this purpose and there exist numerous publications on
this topic. It is not our intention to review these materials here so
the interested reader is referred to recent publications and review
articles on this topic (see e.g. the works451,452,457,465 and references
therein).

12.7.1.3 Ceria in solid oxide fuel cell electrodes


As mentioned above, the requirement for low electrode polarization
resistance points to the use of a MIEC material in order to increase
the active electrode area as much as possible. Among the mixed
conductors with a fluorite structure, ceria-based materials have the
highest ionic and electronic conductivity in reducing conditions
(see Section 12.4.2), and have therefore been thought of as being a
promising alternative anode material. Takahashi et al.466 showed
more than 40 years ago that doped ceria is an excellent hydrogen
electrode at 1000 °C and explained that this was due to the mixed
ionic and electronic conductivity of reduced ceria. However, the
electronic conductivity is still too low, compared to cermets, for
doped ceria to be an effective current collector, which has limited
the practical application of ceria-based materials as single compo-
nents in fuel electrodes. Another drawback of ceria-based materials
is their high stoichiometric expansion coefficient, as discussed in
Section 12.2.2.3, and the associated mechanical problems (see
Section 12.6.5).
The following sections will summarize the use of ceria-based
materials in SOFC electrodes, focusing on the fuel electrode
(anode). It is not our intention to review different fuel electrodes,
but rather to illustrate the importance of ceria for electrodes and
especially to shed some light on the possible mechanisms occurring
when ceria-based materials are used in an electrode. For SOFCs,
there has been much progress during the last decade in using ceria

b1469_Ch-12.indd 735 4/8/2013 12:41:03 PM


b1469 Catalysis by Ceria and Related Materials

736 C. Chatzichristodoulou et al.

as part of the fuel electrode, especially in the form of a nano-sized


material obtained via various ingenious and relatively simple fabrica-
tion methods. We will highlight some of the work that has been
done and give a summary of the ideas that have been brought for-
ward to explain the excellent performance obtained in various fuels
when using nano-sized ceria materials in combination with other
materials and electrocatalysts. Where appropriate, the use of ceria-
based materials in the air electrode will be highlighted and dis-
cussed as well.

12.7.2 Ceria-based electrodes


12.7.2.1 Electrodes for hydrogen-based fuels
Möbius and Rohland467 showed back in 1964 that ceria could work
as an anode in an SOFC fuelled with H2 and CO. Shortly after,
Takahashi et al.466 described Ce0.6Y0.4O1.8 and Ce0.6La0.4O1.8 as being
excellent anode materials for SOFCs. Since then, a number of stud-
ies have been carried out.463,468–471 Gd-doped ceria was investigated as
an SOFC anode and reasonable performance was achieved when H2
was used as the fuel at 1000 °C. Uchida et al.472 investigated Sm-doped
ceria as a potential anode for hydrogen-based fuels. They showed
that the performance could be improved by tailoring the composi-
tion and microstructure of the ceria phase. It was also found that
nanometer-sized Ru catalysts dispersed on the Sm-doped ceria phase
further improved the anodic reactions. Similarly, Primdahl and
Mogensen463 and Primdahl and Liu473 showed that the addition of
only small amounts of Ni (about 1 wt%) to mixed conducting
anodes, including Gd-doped ceria, resulted in a large reduction of
the electrode polarization resistance. A switch of the fuel from
hydrogen with 3% water to deuterium with 3% heavy water affected
the polarization resistance. The low frequency part of the electrode
impedance increased by about 30%. This behavior was explained as
improved adsorption and dissociation of hydrogen by the Ni when
it was dispersed on the mixed conducting phase. Even though the
adsorption part of the explanation may be debated, it indicated that

b1469_Ch-12.indd 736 4/8/2013 12:41:03 PM


b1469 Catalysis by Ceria and Related Materials

Ceria and its Use in Solid Oxide Cells and Oxygen Membranes 737

the hydrogen reaction on the ceria surface is a major rate-limiting


step for SOFC anodes based on Gd-doped CeO2 in H2/H2O mix-
tures at 850 °C, and the addition of a good hydrogen-bond breaker
such as Ni, alleviated this limitation significantly. It is well known,
both experimentally and theoretically, from heterogeneous catalysis
and surface science that Ni and other transition metals (as well as
some noble metals such as Pd and Ru) are excellent catalysts for dis-
sociative adsorption of hydrogen.474–476 However, recent results from
Chueh et al.477 show that ceria is indeed inherently active towards
hydrogen electro-oxidation. They demonstrated that the near-
equilibrium H2 oxidation reaction pathway seems to be dominated
by electrocatalysis at the ceria/gas interface with minimal contribu-
tions from the ceria/metal/gas triple-phase boundaries. Their
results illustrate that the metal-catalysed pathway does not contrib-
ute significantly to the electro-oxidation of hydrogen, at least at low
current densities.
Despite these recent results, the low electronic conductivity of
ceria-based oxides makes it difficult to use them as a single-material
fuel electrode at lower temperatures. The surface reaction kinetics
and lateral electron transport or rate of removal of electrons from
the ceria surface have been reported to be the most likely rate-
limiting processes.478–480 At temperatures below 1000 °C ceria-based
oxides only show promising performance as a component of cer-
mets with Ni or other ceramic electrode materials, or as will be illus-
trated later, when infiltrated as a nano-sized phase in porous
electronic conducting backbones.
Ceria has been used as the ceramic part (or as an addition) in
nickel– or ruthenium–cermet anodes for hydrogen oxidation.472,481–485
Beneficial effects have been reported and interpreted as most likely
being due to the broadening of the three-phase boundary zone.
However, one of the major drawbacks of using ceria in cells with
YSZ-based electrolytes is its chemical reactivity with the YSZ
electrolyte at high temperatures. Sintering of a doped ceria anode
on a YSZ electrolyte at high temperatures (> 1200 °C) results in
the formation of a reaction (diffusion) zone with limited oxide ion
conductivity.203,376,486

b1469_Ch-12.indd 737 4/8/2013 12:41:03 PM


b1469 Catalysis by Ceria and Related Materials

738 C. Chatzichristodoulou et al.

12.7.2.2 Electrodes for hydrocarbon-based fuels


Direct electrochemical oxidation of hydrocarbons in SOFCs
with ceria-based electrodes was claimed in the literature a decade
ago.487–492 There has been some debate in the literature over pre-
cisely what should be called direct oxidation or direct utilization of
hydrocarbons in an SOFC.448 Marina and Mogensen493 and Park et al.488
pointed out that direct electrochemical oxidation of complex
hydrocarbons (see reaction (12.46)) is unlikely to occur in one step.
Even in the case of methane, the reaction should produce eight
electrons and must therefore almost certainly occur in multiple
steps. The open-circuit voltage (OCV) obtained when using hydro-
carbon-based fuels is therefore most likely a mixed potential or one
with a H2/H2O ratio reflecting the amount (or rate) of hydrocarbon
being cracked. The OCV obtained by the authors claiming direct
electrochemical oxidation was much lower than the voltage pre-
dicted by thermodynamics, indicating that the conversion of hydro-
carbons is indeed proceeding by a pathway other than by direct
electrochemical oxidation, i.e. by cracking followed by electrochem-
ical oxidation of the cracking products:

æ
Cx H y ¨
æÆ x C + y H2 (12.47)
2

æ
H2 + O2a ¨
æÆ H2O + 2e- (12.48)

C + 2O 2 - ¨
æ
æÆ CO2 + 4e- (12.49)

However, reactions (12.47) to (12.49) above still represent a very


simplified reaction scheme because the cracking process is usually
highly complex.448 In any case, the anode material in SOFCs must be
able to oxidize the cracking products, and especially the formed
carbon at sufficient rates. The formed carbon (coke) can also be
removed by steam or carbon dioxide:

æ
H2 O + C ¨
æÆ CO + H2 (12.50)

b1469_Ch-12.indd 738 4/8/2013 12:41:03 PM


b1469 Catalysis by Ceria and Related Materials

Ceria and its Use in Solid Oxide Cells and Oxygen Membranes 739

æ
CO2 + C ¨
æÆ 2CO (12.51)

The CO produced can naturally also act as a fuel according to


reaction (12.45). The steam in reaction (12.48) can further reform
the hydrocarbon via steam reforming and assist in providing more
easily oxidized products:

æ
Cx H y + xH2O ¨
æÆ x CO + x + y H2
2 ( ) (12.52)

Work on the use of ceria-based materials as an anode for electro-


chemical oxidation (i.e. without steam reforming) of methane and
other hydrocarbons has been performed by several authors.448,469–471,494
It was initially believed that reduced ceria was a good electrocatalyst
for direct methane oxidation.468,469 This was based on the observa-
tion that in blind tests with a Pt current collector directly on the YSZ
electrolyte, a very high polarization resistance was obtained, and
thus it was concluded that Pt was almost inert for the reactions in
question as also reported by Steele et al.458 The tricky point is, how-
ever, that platinum is a good methane-cracking catalyst. Later,
detailed investigations showed that reduced ceria compounds have
relatively low activity to C–H bond breaking.493,495–497 If an inert gold
current collector was used,493 then the ceria-based anode was almost
inert in the methane, and hence a suitable cracking catalyst had to
be present. The methane conversion observed at OCV and low
anodic overpotentials was more likely due to thermal methane
cracking in the gas phase and on the surfaces in the cell housing.
However, there are other reports with Gd-doped ceria where the
activity for the oxidation of methane was higher than for the oxida-
tion of hydrogen.498 The authors used cone-shaped electrodes in
those oxidation experiments. The apparently contradictory results
could be due to the difficulty of obtaining segregation-free (of
impurities and dopants) surfaces of doped ceria, the right
microstructure of the porous electrodes, or that the interface
between porous Gd-doped ceria and the zirconia-based electrolyte is
different from that obtained on cone-shaped electrodes and

b1469_Ch-12.indd 739 4/8/2013 12:41:03 PM


b1469 Catalysis by Ceria and Related Materials

740 C. Chatzichristodoulou et al.

low-temperature sintered ceria-based electrodes. Different crystal


faces might also be exposed when working with cone-shaped elec-
trodes and porous electrodes,498 which could partly explain some of
the differences.
Ceria is an important component in automotive catalytic
converters499,500 and therefore there is significant information avail-
able about how to improve its catalytic activity and stability. For
example, in a catalytic converter, ceria is present as a mixture with
zirconia.499,500 Without the addition of zirconia, the ceria deactivates
rapidly; in contrast, ceria–zirconia mixtures maintain their activity
for many years under very harsh environments. The mechanism by
which zirconia stabilizes ceria remains poorly understood; however,
it is usually assumed that ceria and zirconia exist as a solid solu-
tion.501 Ceria catalysts can be modified by doping with other rare-
earth oxides, such as Sm2O3, Gd2O3, and La2O3.28,502,503 Some studies
claim that these actually poison the catalytic activity of ceria for
hydrocarbon oxidation in heterogeneous catalysis.504 Replacing
ceria with Sm-doped ceria (SDC) in Cu-ceria-YSZ SOFC anodes
was reported to decrease anode performance.505 However, much of
the difference in performance could actually be due to the differ-
ence observed in the OCV of the two cells, which might be attrib-
uted to a lack of reproducibility in manufacturing cells.488 In
contrast, promising performance has been shown with Ni–SDC
cermet anodes with methane as a fuel.506 Compared to Ni–YSZ, the
cell with the Ni–SDC anode had higher OCV and a lower polariza-
tion resistance. These results illustrate an important point, that
comparing the catalytic performance of ceria-based materials in
heterogeneous catalysis with the electrocatalytic performance of
the same materials in SOFC anodes, based on the materials proper-
ties, is problematic. The ceria-based materials will be in a reduced
state when they operate in a hydrocarbon or hydrogen environ-
ment in an SOFC. The degree of fuel conversion and correspond-
ing changes in gas composition also induce changes in the mixed
transport properties of doped ceria (see Section 12.4.2). This will
naturally affect the surface chemistry and (electro)oxidation
mechanisms.

b1469_Ch-12.indd 740 4/8/2013 12:41:03 PM


b1469 Catalysis by Ceria and Related Materials

Ceria and its Use in Solid Oxide Cells and Oxygen Membranes 741

In summary, interesting results have been obtained using humid-


ified hydrocarbons as a fuel with ceria-based electrodes. The conver-
sion of hydrocarbons is probably not due to direct electrochemical
oxidation of the fuel but rather due to cracking followed by the
electrochemical oxidation of the cracking products. Nevertheless,
the use of ceria seems to be a means for obtaining robust and highly
active electrodes for the oxidation of hydrocarbons. We believe that
there is a real potential for ceria as part of an anode for the conver-
sion of hydrocarbon fuels, because ceria seems to resist carbon pre-
cipitation and is able to oxidize the carbon. The fact that ceria is very
resistant to carbon precipitation is of great importance for maintain-
ing the high reactivity of the fuel electrode during long-term
operation.

12.7.3 Ceria nanoparticles in SOFC electrodes


The need to operate SOFCs at intermediate temperatures has stimu-
lated research into alternative fuel electrode compositions and
ingenious design solutions to replace the established Ni–YSZ cermet
anode, which suffers from various limitations as already mentioned.
One strategy to avoid the use of Ni, or at least the use of Ni as an
electron-conducting backbone phase as in the Ni–YSZ cermet, is to
develop electronically conducting ceramics (or backbone structures)
that are more tolerant to redox cycles and to impurities in the fuel.
Ceramic anodes have been shown to be stable against carbon-fiber
formation in hydrocarbon-based fuels, and, depending on the par-
ticular material used, ceramic anodes can be sulfur tolerant. Some
ceramic materials tested as fuel electrodes have the additional desir-
able properties of being insensitive to oxidation and reduction cycles
and of exhibiting excellent thermal stability.507,508 Unfortunately, the
performance of most ceramic anodes has been modest compared to
that achieved by cells with Ni cermets. Typically, SOFCs with ceramic
anodes must be operated at higher temperatures to produce compa-
rable power densities to Ni cermets. The material requirements for
achieving high performance with ceramic anodes have been summa-
rized and outlined in the review by Atkinson et al.508 A number of

b1469_Ch-12.indd 741 4/8/2013 12:41:03 PM


b1469 Catalysis by Ceria and Related Materials

742 C. Chatzichristodoulou et al.

different ceramic materials have been reported, which meet some of


the requirements, such as the minimum conductivity requirements.
However, it has been more difficult to identify conductive oxides that
also exhibit sufficient ionic conductivity under the reducing condi-
tions in the fuel compartment, sufficient surface reactivity with the
fuel of choice, chemical compatibility with YSZ, and compatible ther-
mal expansion with YSZ.
One very promising solution to overcome these drawbacks with
ceramic materials as fuel electrodes is the infiltration of nanoparti-
cles of ceria-based materials into the porous structures of the fabri-
cated ceramic backbones. The infiltration of various materials has
recently been used in SOFC electrodes (see e.g. the works311,509–512
and references therein for inspiration). Infiltration into composite
fuel electrode backbones is illustrated in Fig. 12.22.
By combining ceria-based materials with even more electrocata-
lytically active materials, and taking advantage of the nanosize of the
materials, promising performances have been shown, as will be illus-
trated in the following sections. The main focus will be on the fuel
electrode due to the intriguing complexity of the interaction of
ceria with various fuels under reducing conditions. However, this is

Figure 12.22 Composite fuel electrode formed by infiltration of an electronically


conducting backbone with ceria-based nanoparticles.

b1469_Ch-12.indd 742 4/8/2013 12:41:04 PM


b1469 Catalysis by Ceria and Related Materials

Ceria and its Use in Solid Oxide Cells and Oxygen Membranes 743

not to say that infiltration of ceria-based materials in air electrodes


is fully understood or less interesting.

12.7.3.1 Ceria-based nanoparticles in conventional SOFC electrodes


Ceria nanoparticles have been shown to improve the performance
of air electrodes in SOFCs considerably. The effect is greatest for
LSM–YSZ-based electrodes.510,511,513,514 The significant decrease in
electrode polarization resistance following infiltration suggests that
the enhanced nano-structured microstructure improves the rate-
limiting step in the oxygen reduction reaction within LSM–YSZ
cathodes. MIEC cathodes such as La1−ySryCo1−zFezO3−x (LSCF), which
already provide sufficient electronic and ionic conductivity, also
benefit from the infiltration of ceria nanoparticles.510,511 It has been
suggested that the improvement is caused by the enhanced ionic
conductivity of the infiltrated ceria phase and possibly from addi-
tional catalytic effects (such as an enhanced dissociative adsorption
of oxygen). This is, however, hardly the case in all reported exam-
ples as undoped ceria (which has low ionic conductivity) has also
been reported to be as effective as doped ceria.514 Therefore,
enhanced ionic conductivity and catalytic properties are most likely
not the complete explanation for the improved performance.515
Impurities (e.g. Si-rich glassy phases) in the triple-phased boundary
region of ceramic composite cathodes and at grain boundaries
between ionic conducting particles hinder ionic transfer. By infil-
trating ceria nanoparticles ionic contact can be improved, which
circumvents the region containing the impurity.513 One more obser-
vation to bear in mind is that infiltration has been shown to signifi-
cantly influence the resulting surface and interface chemistry of the
infiltrated materials. The presence of nitrate ions in the impregna-
tion solution can locally dissolve the backbone material (e.g. LSM),
which may result in a reorganization of the surface.514
Infiltration with ceria or doped ceria nanoparticles into conven-
tional Ni–YSZ-based fuel electrodes has also been shown to improve
the electrode performance of the cell significantly.516,517 Cell perfor-
mance was shown to improve by 20% with a ceria coating and by

b1469_Ch-12.indd 743 4/8/2013 12:41:04 PM


b1469 Catalysis by Ceria and Related Materials

744 C. Chatzichristodoulou et al.

more than 50% with an Sm-doped ceria-modified anode.517 This was


attributed to the high ionic conductivity and electrocatalytic activity
of the doped ceria, leading to extension of the TPB areas (as shown
in Fig. 12.22).
The most significant improvement when infiltrating Ni–YSZ
anodes with ceria-based materials is the apparent sustained sulfur
tolerance when operating with humidified H2 fuel containing H2S at
various concentrations (up to 100 ppm).516,517 The sensitivity of
SOFC anodes to sulfur impurities appears to be related to the stabil-
ity of the sulfur compounds that can be formed with the materials
used in the anodes. Using thermodynamic calculations, the effect
of pO2, pS2, and temperature on the phase behavior of Ni and ceria
was studied over a range relevant to SOFC operating conditions with
sulfur-containing fuels.518,519 Experimental studies on the formation
of Ce–O–S compounds showed good agreement with these thermo-
dynamic predictions. The reaction of Ni with sulfur is predicted to
become more favorable as temperature and hydrogen partial pres-
sure (pH2) decrease. Ceria is increasingly reduced (formation of
Ce3+ and oxygen vacancies) as pO2 decreases and the temperature
increases (see Section 12.3.2) and the stoichiometry of ceria appears
to influence sulfur poisoning.519 Reduced ceria seems to be more
reactive towards sulfur-containing compounds than stoichiometric
ceria. This has been shown when using ceria for high-temperature
gas desulfurization. Reduced ceria is superior to CeO2 in acting as
an H2S sorbent.520 By simplifying a reduced ceria surface as Ce2O3,
reaction (12.53) describes how the reduced ceria surface reacts with
H2S to produce an oxysulfate compound and water:

æ
Ce2O3 + H2S ¨
æÆ Ce2O2S + H2O (12.53)

This reaction is dependent on temperature, pO2, pH2O, and H2S


concentration. However, within the normal operating conditions of
an SOFC anode (e.g. 750 °C and humidified hydrogen as fuel) sul-
fur adsorption on the ceria surfaces is a reversible reaction when
the H2S concentration is below 100 ppm.517,519 This means

b1469_Ch-12.indd 744 4/8/2013 12:41:04 PM


b1469 Catalysis by Ceria and Related Materials

Ceria and its Use in Solid Oxide Cells and Oxygen Membranes 745

that electrochemically active sites of Ni can be protected from


H2S poisoning due to the sulfur adsorption of ceria and doped
ceria. Thus, ceria and doped ceria seem to act as sulfur sorbents in
conventional Ni–YSZ anodes, in addition to extending the TPB
area.

12.7.3.2 Infiltration of ceria-based nanoparticles into YSZ scaffolds


Cu-ceria infiltration into a YSZ backbone structure, with promising
activity for the electrochemical oxidation of both hydrogen and
hydrocarbon-based fuels, has been widely reported in the work of
Gorte and co-workers.459,488,491,505,521,522 The primary goal of using Cu
instead of Ni is the direct utilization of hydrocarbons without steam
reforming, since Cu does not catalyze the formation of carbon in
the way that Ni does. In Cu-based anodes, Cu is believed to provide
only electronic conductivity, while ceria is the mixed conductor and
oxidation catalyst responsible for electrochemical oxidation of the
fuel at the three-phase boundaries. However, due to the low melting
point of Cu (1085 °C), possible sintering of Cu particles during long-
term operation even at relatively low operating temperatures might
become a problem.523
Other groups have recently looked further into adding Cu-ceria
to fuel electrodes, as initially investigated by Gorte and Vohs’ group.
It has been shown that a Cu–CeO2 coating on Ni–YSZ greatly
improved the anode stability in syngas (H2 and CO)524 by facilitating
the water gas shift reaction (12.54) and thereby minimizing the risk
of coke formation (see reversed reaction (12.51)). Cu–CeO2 is an
effective water gas shift catalyst, which can be used in hydrogen pro-
duction by steam reforming.525

æ
CO + H2O ¨
æÆ CO2 + H2 (12.54)

Combining Cu with Co in ceria-based electrodes has shown fur-


ther improvement towards coking, especially with the direct utiliza-
tion of methane.526–528

b1469_Ch-12.indd 745 4/8/2013 12:41:04 PM


b1469 Catalysis by Ceria and Related Materials

746 C. Chatzichristodoulou et al.

12.7.3.3 Ceria-based nanoparticles in electronically


conducting backbones
For perovskite-based fuel electrodes, Mn-doped lanthanum stron-
tium chromite (La0.75Sr0.25Mn0.5Cr0.5O3, LSCM), developed by Tao
and Irvine,529,530 has been infiltrated with ceria-based materials by
other research groups, who reported that an improved electrocata-
lytic activity was achieved when tested with various fuels.531–534 Without
infiltration of a ceria phase into LSCM–YSZ anodes, the oxidation of
methane was considered to be limited by insufficient oxygen ion
conductivity in the lanthanum chromite-based materials.535 Gd-doped
ceria has higher oxide ion conductivity than LSCM, which is also
illustrated by an improved performance of these infiltrated elec-
trodes. The mechanism of methane oxidation in Gd-doped ceria-
infiltrated LSCM anodes in wet CH4 was considered to involve the
partial oxidation of methane by a gas/solid reaction between ceria
and methane-generating CO and H2, followed by electrochemical
oxidation of the products. The added ceria also suppressed coke
formation in these anodes.531
Porous and electronically conducting Nb-doped SrTiO3536 was
infiltrated with Gd-doped ceria by Blennow et al.537 Symmetrical cell
measurements at OCV showed that the electrochemical activity was
maintained or even improved compared to Ni–YSZ fuel electrodes
measured under similar conditions in wet hydrogen. It was shown
that if the microstructure of the ceramic composites is optimized the
activity can be enhanced by several orders of magnitude in some
cases. The electronically conducting Nb-doped SrTiO3 phase basi-
cally acted as a support and current collector for the Gd-doped ceria
electrocatalysts.537 Adding a metal catalyst to the ceramic anode
improves the performance even more. Nb-doped SrTiO3–YSZ com-
posites infiltrated with ceria and Pd have recently been reported by
Gross et al.538 The obtained power density of more than 0.5 W⋅cm−2
with wet hydrogen as fuel at 750 °C is promising, considering that
the cell microstructure was not optimized.
Other electronically conducting SrTiO3 materials have also been
successfully infiltrated with ceria and shown promising performance.

b1469_Ch-12.indd 746 4/8/2013 12:41:04 PM


b1469 Catalysis by Ceria and Related Materials

Ceria and its Use in Solid Oxide Cells and Oxygen Membranes 747

Kurokawa et al.539 fabricated an SOFC anode with high sulfur toler-


ance using Y-doped SrTiO3–YSZ as the porous electrode backbone.
The backbone was infiltrated with nano-sized catalytic ceria and Ru.
The infiltrated electrode had good performance, and the sulfur
tolerance especially was shown to be much higher than conven-
tional Ni–YSZ anodes.539 La-doped SrTiO3, another electronically
conducting ceramic under reducing conditions, has also been
shown to have promising performance as an SOFC anode after infil-
tration of ceria and Cu.540 In humidified hydrogen at 750 °C,
a cell with this specific anode demonstrated power densities above
0.5 W.cm−2.
One important conclusion from the work on infiltrated elec-
tronically conducting ceramic anodes is the great similarities in the
reported data. The data in each case seem to suggest that the ceramic
component simply provides electronic conductivity and that the
(electro)catalytic function is supplied by the materials added sepa-
rately via infiltration. However, this also means that there is certainly
room for performance improvement by (i) further optimization of
the amount and type of infiltrated material, (ii) further optimization
of the microstructure of the porous backbone, and (iii) further
improvement of the current collection (electronic conductivity) of
the porous backbone and support structure.
Highly electronically conducting backbones with subsequent
infiltration of electrocatalytic active materials have been used in an
alternative fuel cell design, based on a ferritic stainless steel sup-
port.541 The metal-supported SOFC has porous and highly electroni-
cally conducting layers, into which electrocatalytically active materials
are infiltrated after sintering. An electrochemical performance
beyond that of a state-of-the-art anode-supported SOFC has been
demonstrated by infiltration of Gd-doped ceria and minor amounts
of Ni into a cermet of the electronically (ferritic stainless steel) and
oxide ionic conducting (YSZ-based) backbone.542,543 Maximum
power densities of more than 1 W⋅cm−2 have been demonstrated in
humidified hydrogen at 650 °C.542 It has also been shown that a
metal-supported SOFC cell can have promising performance by hav-
ing only electronically conducting phases (Nb-doped SrTiO3 in

b1469_Ch-12.indd 747 4/8/2013 12:41:04 PM


b1469 Catalysis by Ceria and Related Materials

748 C. Chatzichristodoulou et al.

combination with ferritic stainless steel) in the anode backbone


structure, into which Gd-doped ceria and minor amounts of Ni are
infiltrated.544 Since ferritic stainless steel has very low activity for the
electrocatalytic oxidation of hydrogen (mainly due to the presence
of a Cr2O3 oxide scale on the steel surface), all electrochemical activ-
ity on the anode side comes from the Gd-doped ceria and Ni
nanoparticles.

12.7.4 Mechanistic aspects of ceria in SOC electrodes


The previous sections described how the infiltration of ceria nano-
particles into working electrodes results in improved (electro)cata-
lytic activity, but more importantly for composite electrodes, an
expansion of the TPB. The question is why does it work so well?
A composite electrode is composed of percolating networks of both
electronically conducting and ionic conducting grains. Infiltrating
with ceria-based materials (as discussed in Section 12.7.3) intro-
duces both ionic and electronic conductivity (in reducing atmos-
pheres) thereby extending the TPB (as illustrated in Fig. 12.22).
It is generally believed that the formation of Ce3+ (especially at
the surfaces) is important for the electrochemical oxidation of vari-
ous fuels, and therefore materials with a low (in absolute value)
enthalpy of reduction should be used. As was thoroughly discussed
in Section 12.3.2, the reducibility of ceria-based materials is facili-
tated by doping at temperatures below 1000 °C. This is also valid for
doped nanocrystalline ceria, because doping has been shown to
enhance reducibility and to improve thermal stability, i.e. hinder
crystallite growth (see Section 12.3.2.3). In this respect, the type of
doping does not seem to have a major influence on the formation
of Ce3+. The formation of Ce3+ will also influence electronic conduc-
tivity in ceria-based materials. In pure and acceptor-doped ceria, an
increase in electronic conductivity with decreasing crystallite size is
generally observed, which is caused by the increase in electronic
defects at the grain boundaries.
Oxide ion conductivity is also of importance for achieving high
performance in SOFC electrodes. Unfortunately, the ionic

b1469_Ch-12.indd 748 4/8/2013 12:41:04 PM


b1469 Catalysis by Ceria and Related Materials

Ceria and its Use in Solid Oxide Cells and Oxygen Membranes 749

conductivity of acceptor-doped ceria tends to decrease with a


decrease in crystallite size (for a detailed discussion see Section
12.4.2.3). Although the transport of oxide ions in acceptor-doped
ceria is reasonably well understood, the mechanism for the exchange
of oxygen between the gas phase and the ceramic is poorly under-
stood. Horita et al.485 concluded that the absorption of oxygen on the
surface of Ce0.8Y0.2O1.9 is the rate-determining step for the oxygen
surface exchange reaction. Lane and Kilner545 suggested that the
surface exchange reaction for Ce0.9Gd0.1O1.95 is controlled by the con-
centration of electronic species. The higher concentration of elec-
tronic species in Gd-doped ceria has also been suggested by
Wachsman and co-workers546 to be the reason for the higher surface
exchange coefficient and lower activation energy, compared to YSZ.
The surface exchange coefficients reported in the literature for
doped ceria vary depending on the pO2 of the gas phase. Y-doped
ceria was reported to show an increase in surface exchange coeffi-
cient k with decreasing pO2 at very low pO2.485 A similar trend was
observed for Gd-doped ceria.205,547 However, Gd-doped ceria has also
been reported to show an increase in k with increasing pO2 at a
higher pO2 range.545 The difference is most likely explained by the
fact that the oxygen exchange reaction depends on the gas species in
the surrounding atmosphere.206,545 The increase of the surface
exchange rate at low pO2 may be related to an increase of the oxygen
nonstoichiometry, implying that the oxygen vacancies play an impor-
tant role in the surface reaction kinetics in this pO2 regime, which
would be important for high performance in fuel electrodes. There
are several reports in the literature that support these statements.
Higher activity in H2O electrolysis than in H2 oxidation has been
observed for various ceria electrodes and interpreted as being due to
the increased surface reduction of ceria during electrolysis (more
reducing potentials), which provided increased surface activity and
electron transport.480,548,549 This again illustrates that the concentra-
tion of Ce3+ and oxygen vacancies are important for high electrocata-
lytic activity and it is consistent with the trend in increased oxygen
surface exchange rates at low pO2 as discussed above. Oxygen
vacancy migration and electronic conduction near the surface have

b1469_Ch-12.indd 749 4/8/2013 12:41:04 PM


b1469 Catalysis by Ceria and Related Materials

750 C. Chatzichristodoulou et al.

also been suggested as significant in the activity of Gd-doped ceria


electrodes.550,551 It has been proposed that the MIEC behavior of ceria
and doped ceria, due to the mixed valance states of Ce (Ce3+/Ce4+)
under the reducing conditions in SOFC fuel electrodes, enables sur-
face activity or electrochemical activity for fuel oxidation or electroly-
sis over areas extending beyond the TPB regions.479,480,552 However,
there is another complicating factor when analyzing surface reaction
rates: the effect of humidity. As discussed in Section 12.3.1, protons
and holes may be created in ceria and doped ceria by the reaction
with water vapor. It has been shown that oxygen exchange rates are
enhanced in the presence of water vapor compared with those deter-
mined in dry conditions (see Yokokawa et al.553 and references
therein). This phenomenon might thus also be relevant for the pos-
sible reaction mechanisms for both fuel electrodes and air electrodes
in SOFCs.553
The surface morphology of the ceria phase must also be consid-
ered when explaining differences in performance between different
electrodes since it seems to depend on which substrate, or crystal-
lographic orientation of the substrate, the ceria phase has been
deposited on.554 The pretreatment and the precursors used for the
infiltration of ceria have been shown to have a strong effect on the
final morphology of the ceria phase555 and consequently perfor-
mance. The high electrochemical activity of infiltrated ceria materi-
als might therefore also be related to the interface energy between
the ceria phase and the backbone phase, creating ceria nanocrystals
with different facets of varied activity.
To further explain the (electro)catalytic properties of ceria, the
morphology and free surface energy of several CeO2 surfaces have
been extensively studied. The mechanism for reduction at the CeO2
(111) and (110) surfaces by H2, based on DFT calculations, has
recently been reported by Chen et al.128 They calculated various
energy barriers for the proposed steps in the reaction pathways for
the reduction by hydrogen on different CeO2 surfaces. The follow-
ing steps can be considered as equivalent to the electrochemical
oxidation of hydrogen in the fuel electrode of an SOFC where the
electrons are “supplied” from the cathode side (at the TPB) and

b1469_Ch-12.indd 750 4/8/2013 12:41:04 PM


b1469 Catalysis by Ceria and Related Materials

Ceria and its Use in Solid Oxide Cells and Oxygen Membranes 751

through the electrolyte (YSZ) as oxide ions via the simplified


reactions:
1
∑∑
VO,YSZ + O2(g) + 2e- Æ OO,YSZ
x

2 (12.55)
x ∑∑ ∑∑
OO,YSZ + VO,YSZ x
Æ OO,ceria + VO,YSZ (12.56)

Chen et al. suggested that the reduction at CeO2 surfaces takes


place via a stepwise mechanism, the first step being the adsorption
and dissociation of H2 with the formation of OH species (with the
x
oxygen being a surface oxygen of ceria, OO,ceria ). It has been sug-
gested that the step which involves the dissociation of H2 generates
more stable OH intermediates. It has an activation barrier of 0.25 eV
and 0.53 eV for the CeO2 (111) and (110) surfaces, respectively.128 It
is agreed in the literature that hydrogen adsorbs exothermally on
ceria surfaces according to reaction (12.57) where two Ce4+ ions are
simultaneously reduced to Ce3+:128,556,557

æ
H2 + 2OxO,ceria + 2CexCe ¨
æÆ 2OO,ceria H ∑ + 2CeCe
¢ (12.57)

Water will form by the joining of neighboring OH groups on the


ceria surface according to reaction (12.58), leaving an oxygen
vacancy behind as water desorbs from the surface:128

æ
2OO,ceria H ∑ ¨
æÆ OxO,ceria + H 2O(g)+ VO,ceria
∑∑
(12.58)

The electrons will be released to the current collector on the


anode side when the cerium goes from Ce3+ back to Ce4+:

¢ ¨
2CeCe æ
Æ 2CexCe + 2e-
æ (12.59)

Combining reactions (12.55) to (12.59) gives the overall reac-


tion for the electrochemical oxidation of hydrogen in an SOFC as
illustrated in Fig. 12.15.
Several authors have found that the reducibility of ceria greatly
depends on the exposed surface (see Section 12.3.2.3). Comparing

b1469_Ch-12.indd 751 4/8/2013 12:41:04 PM


b1469 Catalysis by Ceria and Related Materials

752 C. Chatzichristodoulou et al.

two common surface facets, (110) with (111), suggests that the less
stable (110) surface, compared to (111), is the most energetically
favorable ceria surface for H2 oxidation.128 One reason for this could
be that oxygen vacancies have a lower formation energy on the
(110) surface compared with the (111) surface.126 Similar trends in
surface reactivity have been observed for the oxidation of CO on
various ceria surfaces.123 Experimental studies in heterogeneous
catalysis have shown that CeO2 (100) surfaces are more active than
CeO2 (111) surfaces for CO oxidation.558–60 It has been suggested
that the interaction of CO with ceria surfaces to form CO2 is:123,557

æ
CO + OxO,ceria + 2CexCe ¨
æÆ CO2 + VO,ceria
∑∑
+ 2CeCe
¢ (12.60)

DFT calculations suggest that the mechanism for CO oxidation


on ceria surfaces starts with the CO molecule weakly adsorbing to a
Ce ion at the surface, and then interacting with a neighboring oxide
ion eventually leading to the formation and desorption of CO2.557
As discussed previously, the conversion of hydrocarbons in ceria-
based fuel electrodes most likely proceeds by a pathway other than
direct electrochemical oxidation, i.e., by cracking followed by elec-
trochemical oxidation of the cracking products. This requires an
electrode that is both highly active for the cracking of hydrocarbons
and active for the electrochemical oxidation of the cracking prod-
ucts. The cracking of hydrocarbons also depends on the operating
temperature and the type of hydrocarbon used (the number of
carbon atoms in the chain, OH groups, etc.). It is thus important to
know the pathway for the conversion of hydrocarbons in order to
determine which properties of the electrode should be optimized.
Despite the considerable amount of experimental work mentioned
in the previous sections, the adsorption, dissociation, and oxidation
of even the simplest hydrocarbon (methane) on ceria-based elec-
trodes in SOFCs are processes that are not yet fully understood. DFT
calculations of model systems have recently been undertaken and
reported in the literature, and they provide more insights into the
oxidation of methane on ceria surfaces. For a detailed discussion

b1469_Ch-12.indd 752 4/8/2013 12:41:04 PM


b1469 Catalysis by Ceria and Related Materials

Ceria and its Use in Solid Oxide Cells and Oxygen Membranes 753

regarding these issues refer to e.g. Knapp and Ziegler.557 Their calcu-
lations demonstrate that methane can be dissociated on ceria (111)
surfaces without the adsorption of carbon and the subsequent for-
mation of coke. The abstraction of the first hydrogen atom from the
methane molecule was shown to be a rate-limiting step, with the
creation of a methyl (*CH3) radical in the gas phase and a hydrogen
atom on the surface (on top of a surface oxide ion on the ceria (111)
surface) being the most probable path. The adsorption of a hydro-
gen atom also leads to reduction of one neighboring cerium ion:

æ
CH4 + OxO,ceria + 2CexCe ¨
æÆ * CH3 + OO,ceria H· + CeCe

(12.61)

The following steps in the suggested pathway are the further


abstraction of hydrogen atoms from the hydrocarbon radicals, even-
tually leading to a fully stripped carbon atom with four hydrogen
atoms situated as OH groups on the ceria surface. After all the
hydrogen atoms have been abstracted from the methane molecule
(and radicals), the adsorption of carbon on the ceria (111) surface
leads to the spontaneous desorption of carbon monoxide and the
creation of an oxygen vacancy at the surface.557 This spontaneous
formation of CO from carbon could potentially explain why ceria-
based electrodes show resistance to coking with hydrocarbon-based
fuels. The carbon monoxide formed can further react according to
reaction (12.60) and the four hydrogen atoms on the surface oxy-
gen can react according to reaction (12.48) to form two water mol-
ecules, which desorb from the ceria surface. However, these
vacancy-creating processes require considerable energy, in the range
2.3–2.5 eV for the vacancies formed when water desorbs.556,557 They
would therefore be strongly discouraged in the absence of a vacancy-
filling mechanism. However, it must be emphasized that the genera-
tion of water, CO, and CO2 will be driven by the supply of oxide ions
from the cathode (from the oxygen reduction reaction) to the
anode side via the electrolyte (according to reactions (12.55) and
(12.56)) thereby replenishing the surface oxygen vacancies left
behind by the desorption of water.

b1469_Ch-12.indd 753 4/8/2013 12:41:04 PM


b1469 Catalysis by Ceria and Related Materials

754 C. Chatzichristodoulou et al.

This reaction scheme for methane is simplified but gives a flavor


of the complexity of anodic oxidation reactions. The situation is
more complicated for other hydrocarbons and shows that much
more work is needed in this field in order to find even better elec-
trocatalysts (combinations). More work is also needed before the
vast knowledge of heterogeneous catalysis can be used for
electrocatalysis.

12.7.5 Concluding remarks


Ceria-based materials are of interest as electrode components in
SOFCs due to numerous reports that doped ceria may help decrease
internal electrical resistance by reducing polarization resistance in
both the fuel and the air electrodes.
The literature describes how pure ceria has a higher activity
in heterogeneous catalytic oxidation of hydrocarbons than doped
ceria. The results obtained in the works cited above clearly show that
this is not the case for the anodic conversion of methane or any
other of the commonly used fuels in SOFCs. As discussed in Section
12.7.1.1, the catalytic properties in an oxidizing atmosphere cannot
be directly related to the electrocatalytic properties in a strongly
reducing atmosphere. The composition of ceria as well as the mech-
anism for oxidation vs. electrochemical oxidation of the fuel is dif-
ferent in the two very different atmospheres. At open circuit, no
oxide ions are available from the cathode to react with the fuel.
Evidently, electrochemical oxidation of the fuel at the anode
involves the release of lattice oxygen anions from the ceria phase,
which takes place readily in the reducing atmosphere of the anode
compartment. Accompanying the release of a lattice oxygen anion,
two electrons are liberated. If a highly conductive backbone exists in
the anode, the transfer of electrons from the ceria phase to the adja-
cent electronically conducting phase can occur directly after the
release of lattice oxide ions. Since the liberated electrons are rapidly
transferred to the adjacent electronically conducting phase and
then leave the anode through the current collector connected to
the electrode, lattice oxide ions are steadily released for participa-
tion in the electrochemical oxidation of the fuel.

b1469_Ch-12.indd 754 4/8/2013 12:41:04 PM


b1469 Catalysis by Ceria and Related Materials

Ceria and its Use in Solid Oxide Cells and Oxygen Membranes 755

In general, the factors that can sustain the activity and stability
of ceria-based electrocatalysts in SOFCs seem to be: (i) the flux of
oxide ions through the electrolyte (indirectly from the reduction of
O2 to 2O2− at the cathode); (ii) the release of lattice oxide ions
from the ceria phase (related to the surface exchange coefficient);
(iii) the microstructure of the electrode (electrocatalysts) or the
microstructure of the supporting backbone for infiltrated electro-
catalysts; (iv) current collection or electronic conductivity of the
backbone structure and to some extent in the ceria phase as well.
From the literature it is clear that nano-structured ceria has sev-
eral advantages over similar microcrystalline materials. In SOFC
applications, resistance to impurities such as sulfur, and resistance to
carbon formation when hydrocarbons are used as fuel, have been
reported. The high density of surface defects in nanostructured
materials provides a large number of active sites for various surface
interactions and surface exchange processes. There are clearly inter-
esting features in nano-sized ceria compounds that need further
exploration. The possibility of minimizing pretreatment and water
vapor partial pressure in e.g. the natural gas feed due to lower sus-
ceptibility to coke formation of ceria containing fuel electrodes
(anodes), may simplify the fuel cell system.

Acknowledgments
This work was supported in part by the Danish Agency for Science,
Technology and Innovation through the Strategic Electrochemistry
Research Center (www.serc.dk, contract 2104–06–0011), Catalysis
for Sustainable Energy (www.case.dtu.dk) and the Department of
Energy Conversion and Storage. Financial support from the EU
project METSOFC (FP7-211940) is also gratefully acknowledged.

References
1. Minh, N.Q., Takahashi, T., Science and Technology of Ceramic Fuel Cells,
Elsevier Science B.V., Amsterdam, (1995).
2. Baur, E., Preis, H., Z. Elktrochem. Angew. P, 43 (1937) 727–732.
3. Kröger, F.A., Vink, H.J., Solid State Phys., 3 (1956) 273–301.

b1469_Ch-12.indd 755 4/8/2013 12:41:04 PM


b1469 Catalysis by Ceria and Related Materials

756 C. Chatzichristodoulou et al.

4. Kevane, C.J., Phys. Rev. A, 133 (1964) 1431–1436.


5. Kofstad, P., Hed, A.Z., J. Am. Ceram. Soc., 50 (1967) 681–682.
6. Blumenthal, R.N., Lee, P.W., Panlener, R.J., J. Electrochem. Soc., 118
(1971) 123–129.
7. Steele, B.C.H., Floyd, J.M., Proc. Brit. Ceram. Soc., 19 (1971) 55–76.
8. Faber, J., Seitz, M.A., Mueller, M.H., J. Phys. Chem. Sol., 37 (1976)
903–907.
9. Faber, J., Seitz, M.A., Mueller, M.H., J. Phys. Chem. Sol., 37 (1976)
909–915.
10. JCPDS, International Committee for Diffraction Data, Card #43–1002.
11. Brauer, G., Gingerich, K.A., J. Inorg. Nucl. Chem., 16 (1960) 87–99.
12. Sims, J.R., Blumenthal, R.N., High Temp. Sci., 8 (1976) 99–110.
13. Körner, R., Ricken, M., Nölting, J., Riess, I., J. Solid State Chem., 78
(1989)136–147.
14. Schwab, R.G., Steiner, R.A., Mages, G., Beie, H.J., Thin Solid Films, 207
(1992) 288–293.
15. Chiang, H.W., Blumenthal, R.N., Fournelle, R.A., Solid State Ionics, 66
(1993) 85–95.
16. Mogensen, M., Lindegaard, T., Hansen, U.R., Mogensen, G.,
J. Electrochem. Soc., 141 (1994) 2122–2128.
17. Chatzichristodoulou, C., Hendriksen, P.V., Hagen, A., J. Electrochem.
Soc., 157 (2010) B299–307.
18. Dikmen, S., Shuk, P., Greenblatt, M., Solid State Ionics, 126 (1999)
89–95.
19. Dikmen, S., Shuk, P., Greenblatt, M., Solid State Ionics, 112 (1998)
299–307.
20. Nakamura, A., Solid State Ionics, 181 (2010) 1543–1564.
21. Kim, D.J., J. Am. Ceram. Soc., 72 (1989) 1415–1421.
22. Shannon, R.D., Prewitt, C.T., Acta Crystal. B: Stru., B25 (1969)
925–946.
23. Hong, S.J., Virkar, A.V., J. Am. Ceram. Soc., 78 (1995) 433–439.
24. Mogensen, M., Sammes, N.M., Tompsett, G.A., Solid State Ionics, 129
(2000) 63–94.
25. Etsell, T., Flengas, S., Chem. Rev., 70 (1970) 339–376.
26. Bevan, D.J.M., Summerville, E., Handbook on the Physics and Chemistry
of Rare Earths, North-Holland Physics Publishing, Amsterdam, (1979).

b1469_Ch-12.indd 756 4/8/2013 12:41:05 PM


b1469 Catalysis by Ceria and Related Materials

Ceria and its Use in Solid Oxide Cells and Oxygen Membranes 757

27. Keler, E.K., Godina, N.A., Kalinina, A.M., Zh. Neorgan. Khim. Mosc., 1
(1956) 2557.
28. Eguchi, K., Setoguchi, T., Inoue, T., Arai, H., Solid State Ionics, 52
(1992) 165–172.
29. Christiansen, B.Z., Jacobsen, T., Skaarup, S., Solid State Ionics, 86–88
(1996) 725–731.
30. Ranlov, J., Poulsen, F., Mogensen, M., Solid State Ionics, 61 (1993)
277–279.
31. Sirman, J.D., Waller, D., Kilner, J.A., Electrochem. Soc. Proc., (1997) 1159.
32. Yokokawa, H., J. Phase Equilib., 20 (1999) 258–287.
33. Mogensen, G., Mogensen, M., Thermochim. Acta, 214 (1993) 47–50.
34. Yasuda, I., Hishinuma, M., in Proceedings of the 4th International
Symposium on Solid Oxide Fuel Cells, eds. M. Dokiya, O. Yamamoto, H.
Tagawa, S.C., Singhal, The Electrochemical Society, Pennington, NJ,
(1995), p. 924.
35. Chen, X., Yu, J., Adler, S., Chem. Mater., 17 (2005) 4537–4546.
36. Hendriksen, P.V., Carter, J.D., Mogensen, M., in Proceedings of the 4th
International Symposium on Solid Oxide Fuel Cells, eds. M. Dokiya, O.
Yamamoto, H. Tagawa, S.C., Singhal, The Electrochemical Society,
Pennington, NJ (1995), p. 934.
37. Dawicke, J.W., Blumenthal, R.N., J. Electrochem. Soc., 133 (1986)
904–909.
38. Bevan, D.J.M., Kordis, J., J. Inorg. Nucl. Chem., 26 (1964) 1509–1523.
39. Panlener, R.J., Blumenthal, R.N., Garnier, J.E., J. Phys. Chem. Sol., 36
(1975) 1213–1222.
40. Wagner, C., in Progress in Solid State Chemistry, eds. J.O. McCaldin, G.
Somorjai, Pergamon Press, Oxford, New York, (1975), p. 3.
41. Nowick, A.S., in Diffusion in Crystalline Solids, eds. G.E. Murch, A.S.
Nowick, Academic Press, New York, (1984), p. 143.
42. Schneider, D., Godickemeier, M., Gauckler, L.J., J. Electroceram., 1
(1997) 165–172.
43. Bishop, S.R., Duncan, K.L., Wachsman, E.D., Electrochim. Acta, 54
(2009) 1436–1443.
44. Campserveux, J., Gerdanian, P., J. Chem. Thermodyn., 6 (1974)
795–800.
45. Tuller, H.L., Nowick, A.S., J. Phys. Chem. Sol., 38 (1977) 859–867.

b1469_Ch-12.indd 757 4/8/2013 12:41:05 PM


b1469 Catalysis by Ceria and Related Materials

758 C. Chatzichristodoulou et al.

46. Tuller, H.L., Nowick, A.S., J. Electrochem. Soc., 126 (1979) 209–217.
47. Ling, S., Phys. Rev. B, 49 (1994) 864–880.
48. Bevan, D.J.M., J. Inorg. Nucl. Chem., 1 (1955) 49–56.
49. Zhang, J., Kang, Z.C., Eyring, L., J. Alloy Compd., 192 (1993) 57–63.
50. Kang, Z.C., Eyring, L., J. Alloy Compd., 249 (1997) 206–212.
51. Kümmerle, E.A., Heger, G., J. Solid State Chem., 147 (1999) 485–500.
52. Kümmerle, E.A., Güthoff, F., Schweika, W., Heger, G., J. Solid State
Chem., 153 (2000) 218–230.
53. Zinkevich, M., Djurovic, D., Aldinger, F., Solid State Ionics, 177 (2006)
989–1001.
54. Rezukhina, T.N., Lavrentiev, V.I., Levitskii, V.A., Kuznetsov, F.A., Zh.
Fiz. Khim., 35 (1961) 1367–1369.
55. Markin, T.L., Bones, R.J., Wheeler, V.J., Proc. Brit. Ceram. Soc., 8
(1967) 51.
56. Iwasaki, B., Katsura, T., B. Chem. Soc. Jpn., 44 (1971) 1297–1301.
57. Sorensen, O.T., J. Solid State Chem., 18 (1976) 217–233.
58. Sorensen, O.T., J. Therm. Anal., 13 (1978) 429–437.
59. Touzelin, B., J. Nucl. Mater., 101 (1981) 92–99.
60. Riess, I., Janczikowski, H., Nölting, J., J. Appl. Phys., 61 (1987)
4931–4393.
61. Panhans, M.A., Blumenthal, R.N., Solid State Ionics, 60 (1993)
279–298.
62. Greener, E.H., Wimmer, J.M., Hirthe, W.M., Rare Earth Research III,
Gordan and Breach Inc., New York, (1964).
63. Naik, I.K., Tien, T.Y., J. Phys. Chem. Sol., 39 (1978) 311–315.
64. Kofstad, P., Nonstoichiometry, Diffusion and Electrical Conductivity in
Binary Metal Oxides, Wiley-Interscience, New York, (1972).
65. Koelling, D.D., Boring, A.M., Wood, J.H., Solid State Commun., 47
(1983) 227–232.
66. Wuilloud, E., Delley, B., Schneider, W.D., Baer, Y., J. Magn. Magn.
Mater., 47–48 (1985) 197–199.
67. Allen, J.W., J. Magn. Magn. Mater., 47–48 (1985) 168–174.
68. Park, J.H., Blumenthal, R.N., J. Am. Ceram. Soc., 71 (1988) C462–463.
69. Nakamura, A., J. Nucl. Mater., 201 (1993) 17–26.
70. Duncan, K.L., Wang, Y., Bishop, S.R., Ebrahimi, F., Wachsman, E.D.,
J. Appl. Phys., 101 (2007) 044906.

b1469_Ch-12.indd 758 4/8/2013 12:41:05 PM


b1469 Catalysis by Ceria and Related Materials

Ceria and its Use in Solid Oxide Cells and Oxygen Membranes 759

71. Lindemer, T.B., Calphad, 10 (1986) 129–136.


72. Redlich, O., Kister, A.T., Ind. Eng. Chem., 40 (1948) 345–348.
73. Hillert, M., Jansson, B., J. Am. Ceram. Soc., 69 (1986) 732–734.
74. Yokokawa, H., Sakai, N., Horita, T., Yamaji, K., Xiong, Y., Otake, T.,
Yugami, H., Kawada, T., Mizusaki, J., J. Phase Equilib., 22 (2001) 331–338.
75. Fornasiero, P., Balducci, G., DiMonte, R., Kaspar, J., Sergo, V.,
Gubitosa, G., Ferrero, A., Graziani, M., J. Catal., 164 (1996) 173–183.
76. Vidal, H., Kaspar, J., Pijolat, M., Colon, G., Bernal, S., Cordon, A.,
Perrichon, V., Fally, F., Appl. Catal. B: Environ., 30 (2001) 75–85.
77. Aneggi, E., Boaro, M., de Leitenburg, C., Dolcetti, G., Trovarelli, A.,
J. Alloy Compd., 408 (2006) 1096–1102.
78. Chen, H.T., Chang, J.G., J. Chem. Phys., 132 (2010) 214702.
79. Boaro, M., Desinan, S., Abate, C., Ferluga, M., de Leitenburg, C.,
Trovarelli, A., J. Electrochem. Soc., 158 (2011) P22–29.
80. Chiodelli, G., Flor, G., Scagliotti, M., Solid State Ionics, 91 (1996)
109–121.
81. Tang, Y., Zhang, H., Cui, L., Ouyang, C., Shi, S., Tang, W., Li, H.,
Lee, J., Chen, L., Phys. Rev. B, 82 (2010) 125104.
82. Andersson, D.A., Simak, S.I., Skorodumova, N.V., Abrikosov, I.A.,
Johansson, B., Appl. Phys. Lett., 90 (2007) 031909.
83. Tuller, H.L., Nowick, A.S., J. Electrochem. Soc., 122 (1975) 255–259.
84. Garnier, J.E., Blumenthal, R.N., Panlener, R.J., Sharma, R.K., J. Phys.
Chem. Sol., 37 (1976) 369–378.
85. Park, J.H., Blumenthal, R.N., Panhans, M.A., J. Electrochem. Soc., 135
(1988) 855–859.
86. Wang, S.R., Inaba, H., Tagawa, H., Dokiya, M., Hashimoto, T., Solid
State Ionics, 107 (1998) 73–79.
87. Kobayashi, T., Wang, S.R., Dokiya, M., Tagawa, H., Hashimoto, T.,
Solid State Ionics, 126 (1999) 349–357.
88. Djurovic, D., Zinkevich, M., Aldinger, F., Solid State Ionics, 179 (2008)
1902–1911.
89. Osaka, M., J. Alloy Compd., 475 (2009) L31–33.
90. Wang, S., Inaba, H., Tagawa, H., Hashimoto, T., J. Electrochem. Soc., 144
(1997) 4076–4080.
91. Mogensen, M., in Catalysis by Ceria and Related Materials, ed.,
A. Trovarelli, Imperial College Press, London, (2002), p. 453.

b1469_Ch-12.indd 759 4/8/2013 12:41:05 PM


b1469 Catalysis by Ceria and Related Materials

760 C. Chatzichristodoulou et al.

92. Vanderlaan, G., Fuggle, J.C., Vandijk, M.P., Burggraaf, A.J., Esteva, J.M.,
Karnatak, R., J. Phys. Chem. Sol., 47 (1986) 413–416.
93. Martinez-Arias, A., Hungria, A.B., Fernandez-Garcia, M., Iglesias-Juez, A.,
Conesa, J.C., Mather, G.C., Munuera, G., J. Power Sources, 151 (2005)
43–51.
94. Fernandez-Garcia, M., Wang, X.Q., Belver, C., Iglesias-Juez, A.,
Hanson, J.C., Rodriguez, J.A., Chem. Mater., 17 (2005) 4181–4193.
95. Borchert, H., Frolova, Y.V., Kaichev, V.V., Prosvirin, I.P., Alikina, G.M.,
Lukashevich, A.I., Zaikovskii, V.I., Moroz, E.M., Trukhan, S.N.,
Ivanov, V.P., Paukshtis, E.A., Bukhtiyarov, V.I., Sadykov, V.A., J. Phys.
Chem. B, 109 (2005) 5728–5738.
96. Lin, X., Yan, S., Zhu, L., Li, L., Su, W., Acta Chim. Sinica, 67 (2009)
1389–1394.
97. Knauth, P., Tuller, H.L., J. Eur. Ceram. Soc., 19 (1999) 831–836.
98. Fagg, D.P., Marozau, I.P., Shaula, A.L., Kharton, V.V., Frade, J.R.,
J. Solid State Chem., 179 (2006) 3347–3356.
99. Chatzichristodoulou, C., Hendriksen, P.V., J. Electrochem. Soc., 157
(2010) B481–489.
100. Bishop, S.R., Stefanik, T.S., Tuller, H.L., PCCP, 13 (2011)
10165–10173.
101. Wang, X.Q., Hanson, J.C., Rodriguez, J.A., Belver, C., Fernandez-
Garcia, M., J. Chem. Phys., 122 (2005) 154711.
102. Fagg, D.P., Frade, J.R., Kharton, V.V., Marozau, I.P., J. Solid State Chem.,
179 (2006) 1469–1477.
103. Hungria, A.B., Martinez-Arias, A., Fernandez-Garcia, M., Iglesias-Juez,
A., Guerrero-Ruiz, A., Calvino, J.J., Conesa, J.C., Soria, J., Chem. Mater.,
15 (2003) 4309–4316.
104. Sakai, N., Yamaji, K., Horita, T., Yokokawa, H., Hirata, Y., Sameshima, S.,
Nigara, Y., Mizusaki, J., Solid State Ionics, 125 (1999) 325–331.
105. Sakai, N., Yamaji, K., Xiong, Y.P., Kishimoto, H., Horita, T.,
Yokokawa, H., J. Electroceram., 13 (2004) 677–682.
106. Yokokawa, H., Horita, T., Sakai, N., Yamaji, K., Brito, M.E., Xiong, Y.,
Kishimoto, H., Solid State Ionics, 177 (2006) 1705–1714.
107. Kim, S., Merkle, R., Maier, J., Solid State Ionics, 161 (2003) 113–119.
108. Yao, H.C., Yao, Y.F.Y., J. Catal., 86 (1984) 254–265.
109. Johnson, M.F.L., Mooi, J., J. Catal., 103 (1987) 502–505.

b1469_Ch-12.indd 760 4/8/2013 12:41:05 PM


b1469 Catalysis by Ceria and Related Materials

Ceria and its Use in Solid Oxide Cells and Oxygen Membranes 761

110. Bernal, S., Calvino, J.J., Cifredo, G.A., Gatica, J.M., Omil, J.A.P.,
Pintado, J.M., J. Chem. Soc. Faraday T., 89 (1993) 3499–3505.
111. Zimmer, P., Tschöpe, A., Birringer, R., J. Catal., 205 (2002) 339–345.
112. Tschöpe, A., J. Electroceram., 14 (2005) 5–23.
113. Trudeau, M.L., Tschöpe, A., Ying, J.Y., Surf. Interface Anal., 23 (1995)
219–226.
114. Tschöpe, A., Birringer, R., Nanostruct. Mater., 9 (1997) 591–594.
115. Porat, O., Tuller, H.L., Lavik, E.B., Chiang, Y.M., Mat. Res. S. C, 457
(1997) 99–103.
116. Janvier, C., Pijolat, M., Valdivieso, F., Soustelle, M., Solid State Ionics,
127 (2000) 207–222.
117. Chiang, Y.M., Lavik, E.B., Kosacki, I., Tuller, H.L., Ying, J.Y., J. Electroceram.,
1 (1997) 7–14.
118. Hwang, J.H., Mason, T.O., Z. Phys. Chem., 207 (1998) 21–38.
119. Kim, S., Fleig, J., Maier, J., Elec. Soc. S., 2001 (2002) 208–217.
120. Kim, S., Merkle, R., Maier, J., Surf. Sci., 549 (2004) 196–202.
121. Suzuki, T., Kosacki, I., Anderson, H.U., Colomban, P., J. Am. Ceram.
Soc., 84 (2001) 2007–2014.
122. Kosacki, I., Suzuki, T., Anderson, H.U., Colomban, P., Solid State Ionics,
149 (2002) 99–105.
123. Sayle, T.X.T., Parker, S.C., Catlow, C.R.A., Surf. Sci., 316 (1994)
329–336.
124. Conesa, J.C., Surf. Sci., 339 (1995) 337–352.
125. Yang, Z.X., Woo, T.K., Baudin, M., Hermansson, K., J. Chem. Phys., 120
(2004) 7741–7749.
126. Fabris, S., Vicario, G., Balducci, G., de Gironcoli, S., Baroni, S., J. Phys.
Chem. B, 109 (2005) 22860–22867.
127. Nolan, M., Fearon, J.E., Watson, G.W., Solid State Ionics, 177 (2006)
3069–3074.
128. Chen, H.T., Choi, Y.M., Liu, M.L., Lin, M.C., Chemphyschem, 8 (2007)
849–855.
129. Bernal, S., Blanco, G., Cauqui, M.A., Corchado, P., Pintado, J.M.,
Rodriguez-Izquierdo, J.M., Chem. Commun., 16 (1997) 1545–1546.
130. Mista, W., Rayment, T., Hanuza, J., Macalik, L., Mater. Sci., 22 (2004)
153–170.
131. Zhao, M., Shen, M., Wang, J., J. Catal., 248 (2007) 258–267.

b1469_Ch-12.indd 761 4/8/2013 12:41:05 PM


b1469 Catalysis by Ceria and Related Materials

762 C. Chatzichristodoulou et al.

132. Hernandez, J.C., Hungria, A.B., Perez-Omil, J.A., Trasobares, S.,


Bernal, S., Midgley, P.A., Alavi, A., Calvino, J.J., J. Phys. Chem. C, 111
(2007) 9001–9004.
133. Malecka, M.A., Kepinski, L., Mista, W., Appl. Catal. B: Environ., 74
(2007) 290–298.
134. Malecka, M.A., Kepinski, L., Mista, W., J. Alloy Compd., 451 (2008)
567–570.
135. Bernal, S., Blanco, G., Calvino, J.J., Hernandez, J.C., Perez-Omil, J.A.,
Pintado, J.M., Yeste, M.R., J. Alloy Compd., 451 (2008) 521–525.
136. Mista, W., Malecka, M.A., Kepinski, L., Appl. Catal. A: Gen., 368 (2009)
71–78.
137. Lei, Y.Y., Ito, Y., Browning, N.D., Mazanec, T.J., J. Am. Ceram. Soc., 85
(2002) 2359–2363.
138. Yan, M.F., Cannon, R.M., Bowen, H.K., J. Appl. Phys., 54 (1983)
764–778.
139. Luo, M.F., Yan, Z.L., Jin, L.Y., He, M., J. Phys. Chem. B, 110 (2006)
13068–13071.
140. Blom, D.A., Chiang, Y.M., Mat. Res. S. C, 458 (1997) 127–132.
141. Knauth, P., Tuller, H.L., Solid State Ionics, 136 (2000) 1215–1224.
142. Nörenberg, H., Briggs, G.A.D., Phys. Rev. Lett., 79 (1997) 4222–4225.
143. Namai, Y., Fukui, K.I., Iwasawa, Y., Catal. Today, 85 (2003) 79–91.
144. Esch, F., Fabris, S., Zhou, L., Montini, T., Africh, C., Fornasiero, P.,
Comelli, G., Rosei, R., Science, 309 (2005) 752–755.
145. Crozier, P.A., Wang, R., Sharma, R., Ultramicroscopy, 108 (2008)
1432–1440.
146. Noddack, W., Walch, H., Z. Elektrochem., 63 (1959) 269–274.
147. Rudolph, J., Z. Naturforsch. Pt. A, 14 (1959) 727–737.
148. Blumenthal, R.N., Pinz, B.A., J. Appl. Phys., 38 (1967) 2376–2383.
149. Vinokurov, I.V., Zonn, Z.N., Ioffe, V.A., Sov. Phys. Sol. State, 9 (1968)
2659–2663.
150. Tschöpe, A., Sommer, E., Birringer, R., Solid State Ionics, 139 (2001)
255–265.
151. Blumenthal, R.N., Panlener, R.J., J. Phys. Chem. Sol., 31 (1970)
1190–1192.
152. Blumenthal, R.N., Sharma, R.K., J. Solid State Chem., 13 (1975)
360–364.

b1469_Ch-12.indd 762 4/8/2013 12:41:05 PM


b1469 Catalysis by Ceria and Related Materials

Ceria and its Use in Solid Oxide Cells and Oxygen Membranes 763

153. Blumenthal, R.N., Hofmaier, R.L., J. Electrochem. Soc., 121 (1974)


126–131.
154. Boureau, G., Masmoudi, O., Tetot, R., Solid State Commun., 79 (1991)
299–302.
155. Brugner, F.S., Blumenthal, R.N., J. Am. Ceram. Soc., 54 (1971) 57.
156. Baker, E.H., Iqbal, M., Knox, B.E., J. Mater. Sci., 12 (1977) 305–310.
157. Pfau, A., Schierbaum, K.D., Surf. Sci., 321 (1994) 71–80.
158. Tschöpe, A., Solid State Ionics, 139 (2001) 267–280.
159. Kim, S., Maier, J., J. Electrochem. Soc., 149 (2002) J73–83.
160. Rupp, J.L.M., Gauckler, L.J., Solid State Ionics, 177 (2006) 2513–2518.
161. Mansilla, C., Holgado, J.P., Espinos, J.P., Gonzalez-Elipe, A.R.,
Yubero, F., Surf. Coat. Tech., 202 (2007) 1256–1261.
162. Litzelman, S.J., Tuller, H.L., Solid State Ionics, 180 (2009) 1190–1197.
163. Litzelman, S.J., De Souza, R.A., Butz, B., Tuller, H.L., Martin, M.,
Gerthsen, D., J. Electroceram., 22 (2009) 405–415.
164. Kim, S., Lee, J.S., Mitterbauer, C., Ramasse, Q.M., Sarahan, M.C.,
Browning, N.D., Park, H.J., Chem. Mater., 21 (2009) 1182–1186.
165. Wang, D.Y., Nowick, A.S., Am. Ceram. Soc. Bull., 59 (1980) 333.
166. Wang, D.Y., Nowick, A.S., J. Solid State Chem., 35 (1980) 325–333.
167. Eladham, K., Hammou, A., Solid State Ionics, 9–10 (1983) 905–912.
168. Gerhardt, R., Nowick, A.S., J. Am. Ceram. Soc., 69 (1986) 641–646.
169. Gerhardt, R., Nowick, A.S., Mochel, M.E., Dumler, I., J. Am. Ceram.
Soc., 69 (1986) 647–651.
170. Tanaka, J., Baumard, J.F., Abelard, P., J. Am. Ceram. Soc., 70 (1987)
637–643.
171. Christie, G.M., Berkel, F.P.F.V., Solid State Ionics, 83 (1996) 17–27.
172. Herle, J.V., Horita, T., Kawada, T., Sakai, N., Yokokawa, H., Dokiya, M.,
J. Eur. Ceram. Soc., 16 (1996) 961–973.
173. Ralph, J.M., Przydatek, J., Kilner, J.A., Seguelong, T., Ber. Bunsen Phys.
Chem., 101 (1997) 1403–1407.
174. Chiang, Y.M., Lavik, E.B., Blom, D.A., Nanostruct. Mater., 9 (1997)
633–642.
175. Hong, S.J., Mehta, K., Virkar, A.V., J. Electrochem. Soc., 145 (1998)
638–647.
176. Hwang, J.H., McLachlan, D.S., Mason, T.O., J. Electroceram., 3 (1999)
7–16.

b1469_Ch-12.indd 763 4/8/2013 12:41:05 PM


b1469 Catalysis by Ceria and Related Materials

764 C. Chatzichristodoulou et al.

177. Li, G.S., Mao, Y.C., Li, L.P., Feng, S.H., Wang, M.Q., Yao, X., Chem.
Mater., 11 (1999) 1259–1266.
178. Tian, C.Y., Chan, S.W., Solid State Ionics, 134 (2000) 89–102.
179. Shemilt, J.E., Stanway, C.L., Williams, H.M., Solid State Ionics, 134
(2000) 111–117.
180. Kleinlogel, C.M., Gauckler, L.J., J. Electroceram., 5 (2000) 231–243.
181. Zhan, Z.L., Wen, T.L., Tu, H.Y., Lu, Z.Y., J. Electrochem. Soc., 148 (2001)
A427–432.
182. Li, L.P., Lin, X.M., Li, G.S., Inomata, H., J. Mater. Res., 16 (2001)
3207–3213.
183. Zhou, X.D., Huebner, W., Kosacki, I., Anderson, H.U., J. Am. Ceram.
Soc., 85 (2002) 1757–1762.
184. Zhang, T.S., Zeng, Z.Q., Huang, H.T., Hing, P., Kilner, J., Mater. Lett.,
57 (2002) 124–129.
185. Zhang, T.S., Hing, P., Huang, H.T., Kilner, J., Solid State Ionics, 148
(2002) 567–573.
186. Zhang, T.S., Ma, J., Huang, H.T., Hing, P., Xia, Z.T., Chan, S.H.,
Kilner, J.A., Solid State Sci., 5 (2003) 1505–1511.
187. Zhang, T.S., Kong, L.B., Zeng, Z.Q., Huang, H.T., Hing, P., Xia, Z.T.,
Kilner, J., J. Solid State Electr., 7 (2003) 348–354.
188. Jung, G.B., Huang, T.J., J. Mater. Sci., 38 (2003) 2461–2468.
189. Perez-Coll, D., Nunez, P., Frade, J.R., Abrantes, J.C.C., Electrochim.
Acta, 48 (2003) 1551–1557.
190. Guo, X., Waser, R., Prog. Mater. Sci., 51 (2006) 151–210.
191. Guo, X., Sigle, W., Maier, J., J. Am. Ceram. Soc., 86 (2003) 77–87.
192. Blumenthal, R.N., Brugner, F.S., Garnier, J.E., J. Electrochem. Soc., 120
(1973) 1230–1237.
193. Kudo, T., Obayashi, H., J. Electrochem. Soc., 123 (1976) 415–419.
194. Wang, D.Y., Park, D.S., Griffith, J., Nowick, A.S., Solid State Ionics, 2
(1981) 95–105.
195. Gerhardt-Anderson, R., Nowick, A.S., Solid State Ionics, 5 (1981) 547–550.
196. Arai, H., Kunisaki, T., Shimizu, Y., Seiyama, T., Solid State Ionics, 20
(1986) 241–248.
197. Yahiro, H., Eguchi, K., Arai, H., Solid State Ionics, 21 (1986) 37–47.
198. Yahiro, H., Ohuchi, T., Eguchi, K., Arai, H., J. Mater. Sci., 23 (1988)
1036–1041.

b1469_Ch-12.indd 764 4/8/2013 12:41:05 PM


b1469 Catalysis by Ceria and Related Materials

Ceria and its Use in Solid Oxide Cells and Oxygen Membranes 765

199. Yahiro, H., Eguchi, Y., Eguchi, K., Arai, H., J. Appl. Electrochem., 18
(1988) 527–531.
200. Yahiro, H., Eguchi, K., Arai, H., Solid State Ionics, 36 (1989) 71–75.
201. Faber, J., Geoffroy, C., Roux, A., Sylvestre, A., Abelard, P., Appl. Phys.
A: Mater., 49 (1989) 225–232.
202. Balazs, G.B., Glass, R.S., Solid State Ionics, 76 (1995) 155–162.
203. Tsoga, A., Naoumidis, A., Stover, D., Solid State Ionics, 135 (2000)
403–409.
204. Wang, S.R., Kobayashi, T., Dokiya, M., Hashimoto, T., J. Electrochem.
Soc., 147 (2000) 3606–3609.
205. Katsuki, M., Wang, S.R., Yasumoto, K., Dokiya, M., Solid State Ionics,
154 (2002) 589–595.
206. Yashiro, K., Onuma, S., Kaimai, A., Nigara, Y., Kawada, T., Mizusaki, J.,
Kawamura, K., Horita, T., Yokokawa, H., Solid State Ionics, 152 (2002)
469–476.
207. Omar, S., Wachsman, E.D., Nino, J.C., Solid State Ionics, 178 (2008)
1890–1897.
208. Omar, S., Wachsman, E.D., Jones, J.L., Nino, J.C., J. Am. Ceram. Soc.,
92 (2009) 2674–2681.
209. Ding, D., Liu, B., Gong, M., Liu, X., Xia, C., Electrochim. Acta, 55
(2010) 4529–4535.
210. Li, B., Liu, Y., Wei, X., Pan, W., J. Power Sources, 195 (2010) 969–976.
211. Yashima, M., Takizawa, T., J. Phys. Chem. C, 114 (2010) 2385–2392.
212. Yang, F., Zhao, X., Xiao, P., J. Power Sources, 196 (2011) 4943–4949.
213. Yao, H., Zhang, Y., Liu, J., Li, Y., Wang, J., Li, Z., Mater. Res. Bull., 46
(2011) 75–80.
214. Kilner, J.A., Waters, C.D., Solid State Ionics, 6 (1982) 253–259.
215. Kilner, J.A., Brook, R.J., Solid State Ionics, 6 (1982) 237–252.
216. Steele, B.C.H., Solid State Ionics, 129 (2000) 95–110.
217. Kilner, J.A., Solid State Ionics, 8 (1983) 201–207.
218. Wang, F., Wan, B., Cheng, S., J. Solid State Electr., 9 (2005) 168–173.
219. Omar, S., Wachsman, E.D., Nino, J.C., Solid State Ionics, 177 (2006)
3199–3203.
220. Xu, H., Yan, H., Chen, Z., Solid State Sci., 10 (2008) 1179–1184.
221. Kumar, V.P., Reddy, Y.S., Kistaiah, P., Prasad, G., Reddy, C.V., Mater.
Chem. Phys., 112 (2008) 711–718.

b1469_Ch-12.indd 765 4/8/2013 12:41:05 PM


b1469 Catalysis by Ceria and Related Materials

766 C. Chatzichristodoulou et al.

222. Guan, X., Zhou, H., Wang, Y., Zhang, J., J. Alloy Compd., 464 (2008)
310–316.
223. Liu, Y., Li, B., Wei, X., Pan, W., J. Am. Ceram. Soc., 91 (2008)
3926–3930.
224. Zheng, Y., Wu, L., Gu, H., Gao, L., Chen, H., Guo, L., J. Alloy Compd.,
486 (2009) 586–589.
225. Yamamura, H., Katoh, E., Ichikawa, M., Kakinuma, K., Mori, T.,
Haneda, H., Electrochemistry, 68 (2000) 455–459.
226. Yamamura, H., Katoh, E., Kakinuma, K., J. Ceram. Soc. Jpn., 110 (2002)
1021–1024.
227. Butler, V., Catlow, C.R.A., Fender, B.E.F., Harding, J.H., Solid State
Ionics, 8 (1983) 109–113.
228. Minervini, L., Zacate, M.O., Grimes, R.W., Solid State Ionics, 116 (1999)
339–349.
229. Andersson, D.A., Simak, S.I., Skorodumova, N.V., Abrikosov, I.A.,
Johansson, B., P. Natl. Acad. Sci. USA, 103 (2006) 3518–3521.
230. Wei, X., Pan, W., Cheng, L., Li, B., Solid State Ionics, 180 (2009) 13–17.
231. Frayret, C., Villesuzanne, A., Pouchard, M., Mauvy, F., Bassat, J.,
Grenier, J., J. Phys. Chem. C, 114 (2010) 19062–19076.
232. Dholabhai, P.P., Adams, J.B., Crozier, P., Sharma, R., J. Chem. Phys., 132
(2010) 094104.
233. Fuda, K., Kishio, K., Yamauchi, S., Fueki, K., Onoda, Y., J. Phys. Chem.
Sol., 45 (1984) 1253–1257.
234. Navrotsky, A., J. Mater. Chem., 20 (2010) 10577–10587.
235. Avila-Paredes, H.J., Shvareva, T., Chen, W., Navrotsky, A., Kim, S.,
PCCP, 11 (2009) 8580–8585.
236. Li, H.B., Xia, C.R., Zhu, M.H., Zhou, Z.X., Meng, G.Y., Acta Mater., 54
(2006) 721–727.
237. Esposito, V., Traversa, E., J. Am. Ceram. Soc., 91 (2008) 1037–1051.
238. Mori, T., Drennan, J., Wang, Y.R., Lee, J.H., Li, J.G., Ikegami, T.,
J. Electrochem. Soc., 150 (2003) A665–673.
239. Mori, T., Wang, Y.R., Drennan, J., Auchterlonie, G., Li, J.G., Ikegami,
T., Solid State Ionics, 175 (2004) 641–649.
240. Ou, D.R., Mori, T., Ye, F., Takahashi, M., Zou, J., Drennan, J., Acta
Mater., 54 (2006) 3737–3746.
241. Ye, F., Mori, T., Ou, D.R., Takahashi, M., Zou, J., Drennan, J.,
J. Electrochem. Soc., 154 (2007) B180–185.

b1469_Ch-12.indd 766 4/8/2013 12:41:05 PM


b1469 Catalysis by Ceria and Related Materials

Ceria and its Use in Solid Oxide Cells and Oxygen Membranes 767

242. Ou, D.R., Mori, T., Ye, F., Zou, J., Auchterlonie, G., Drennan, J.,
J. Electrochem. Soc., 154 (2007) B616–622.
243. Mori, T., Kobayashi, T., Wang, Y., Drennan, J., Nishimura, T., Li, J.G.,
Kobayashi, H., J. Am. Ceram. Soc., 88 (2005) 1981–1984.
244. Ye, F., Mori, T., Ou, D.R., Zou, J., Auchterlonie, G., Drennan, J.,
J. Appl. Phys., 101 (2007) 113528.
245. Li, Z., Mori, T., Auchterlonie, G.J., Zou, J., Drennan, J., Appl. Phys.
Lett., 98 (2011) 093104.
246. Deguchi, H., Yoshida, H., Inagaki, T., Horiuchi, M., Solid State Ionics,
176 (2005) 1817–1825.
247. McCullough, J., Britton, J., J. Am. Chem. Soc., 74 (1952) 5225–5227.
248. Ralph, J.M., Kilner, J.A., in Proceedings of the 2nd European SOFC Forum,
ed. B. Thorstensen, European Fuel Cell Forum Oberrohrdorf,
Switzerland, (1996), p. 773.
249. Zhang, T., Ma, J., Kong, L., Chan, S., Kilner, J., Solid State Ionics, 170
(2004) 209–217.
250. Levy, M., Fouletier, J., Solid State Ionics, 12 (1984) 467–472.
251. Xiong, Y., Yamaji, K., Kishimoto, H., Brito, M.E., Horita, T.,
Yokokawa, H., J. Electrochem. Soc., 155 (2008) B1300–1306.
252. Xiong, Y.P., Yamaji, K., Horita, T., Sakai, N., Yokokawa, H., J. Electrochem.
Soc., 151 (2004) A407–412.
253. Shimonosono, T., Hirata, Y., Ehira, Y., Sameshima, S., Horita, T.,
Yokokawa, H., Solid State Ionics, 174 (2004) 27–33.
254. Perez-Coll, D., Aguadero, A., Nunez, P., Frade, J.R., Int. J. Hydrogen
Energ., 35 (2010) 11448–11455.
255. Navarro, I., Marques, F., Frade, J., J. Electrochem. Soc., 144 (1997)
267–273.
256. Matsui, T., Minoru, I., Mineshige, A., Ogumi, Z., Solid State Ionics, 176
(2005) 647–654.
257. Gorelov, V.P., Balakireva, V.B., Yaroslavtsev, I.Y., Kazantsev, V.A.,
Vaganov, E.G., Russ. J. Electrochem., 43 (2007) 888–893.
258. Takasu, Y., Sugino, T., Matsuda, Y., J. Appl. Electrochem., 14 (1984)
79–81.
259. Nauer, M., Ftikos, C., Steele, B., J. Eur. Ceram. Soc., 14 (1994) 493–499.
260. Huang, W., Shuk, P., Greenblatt, M., Solid State Ionics, 113 (1998)
305–310.
261. Lübke, S., Wiemhöfer, H.D., Solid State Ionics, 117 (1999) 229–243.

b1469_Ch-12.indd 767 4/8/2013 12:41:05 PM


b1469 Catalysis by Ceria and Related Materials

768 C. Chatzichristodoulou et al.

262. Shuk, P., Greenblatt, M., Croft, M., Chem. Mater., 11 (1999) 473–479.
263. Shuk, P., Greenblatt, M., Solid State Ionics, 116 (1999) 217–223.
264. Kharton, V., Viskup, A., Figueiredo, F., Naumovich, E., Yaremchenko, A.,
Marques, F., Electrochim. Acta, 46 (2001) 2879–2889.
265. Stefanik, T., Tuller, H., Solid State Ionics, 756 (2003) 163–168.
266. Qi, X., Lin, Y., Holt, C., Swartz, S., J. Mater. Sci., 38 (2003)
1073–1079.
267. Fagg, D., Kharton, V., Shaula, A., Marozau, I., Frade, J., Solid State
Ionics, 176 (2005) 1723–1730.
268. Fagg, D.P., Garcia-Martin, S., Kharton, V.V., Frade, J.R., Chem. Mater.,
21 (2009) 381–391.
269. Chatzichristodoulou, C., Park, W., Kim, H., Hendriksen, P.V., Yoo, H.,
PCCP, 12 (2010) 9637–9649.
270. Schmale, K., Gruenebaum, M., Janssen, M., Baumann, S., Schulze-
Kueppers, F., Wiemhoefer, H., Phys. Stat. Sol. B, 248 (2011) 314–322.
271. Balaguer, M., Solis, C., Serra, J.M., Chem. Mater., 23 (2011) 2333–2343.
272. Chatzichristodoulou, C., Hendriksen, P.V., PCCP, 13 (2011)
21558–21572.
273. Park, W., Yang, I., Yoo, H.I., ECS Transactions, 13 (2008) 327–336.
274. Lee, H.B., Prinz, F.B., Cai, W., Acta Mater., 58 (2010) 2197–2206.
275. Göbel, M.C., Gregori, G., Maier, J., PCCP, 13 (2011) 10940–10945.
276. Yeh, T.C., Perry, N.H., Mason, T.O., J. Am. Ceram. Soc., 94 (2011)
1073–1078.
277. Bonanos, N., Lilley, E., J. Phys. Chem. Sol., 42 (1981) 943–952.
278. Rodrigo, K., Heiroth, S., Lundberg, M., Bonanos, N., Kant, K.M.,
Pryds, N., Kuhn, L.T., Esposito, V., Linderoth, S., Schou, J., Lippert, T.,
Appl. Phys. A: Mater., 101 (2010) 601–607.
279. Suzuki, T., Kosacki, I., Anderson, H.U., Solid State Ionics, 151 (2002)
111–121.
280. Choi, S.H., Hwang, C.S., Lee, H., Kim, J., J. Electrochem. Soc., 156
(2009) B381–385.
281. Swanson, M., Tangtrakarn, N., Sunder, M., Moran, P.D., Solid State
Ionics, 181 (2010) 379–385.
282. Chiodelli, G., Malavasi, L., Massarotti, V., Mustarelli, P., Quartarone, E.,
Solid State Ionics, 176 (2005) 1505–1512.
283. Joo, J.H., Choi, G.M., J. Eur. Ceram. Soc., 27 (2007) 4273–4277.

b1469_Ch-12.indd 768 4/8/2013 12:41:05 PM


b1469 Catalysis by Ceria and Related Materials

Ceria and its Use in Solid Oxide Cells and Oxygen Membranes 769

284. Huang, H., Gur, T.M., Saito, Y., Prinz, F., Appl. Phys. Lett., 89 (2006)
143107.
285. Beckel, D., Bieberle-Huetter, A., Harvey, A., Infortuna, A.,
Muecke, U.P., Prestat, M., Rupp, J.L.M., Gauckler, L.J., J. Power Sources,
173 (2007) 325–345.
286. Karthikeyan, A., Tsuchiya, M., Chang, C., Ramanathan, S., Appl. Phys.
Lett., 90 (2007) 263108.
287. Lee, W., Lee, M., Kim, Y., Prinz, F.B., Nanotechnology, 20 (2009) 445706.
288. Shirpour, M., Gregori, G., Merkle, R., Maier, J., PCCP, 13 (2011)
937–940.
289. Minh, N., J. Am. Ceram. Soc., 76 (1993) 563–588.
290. Appleby, A., Energy, 21 (1996) 521–653.
291. Ormerod, R., Chem. Soc. Rev., 32 (2003) 17–28.
292. Haile, S., Acta Mater., 51 (2003) 5981–6000.
293. Singhal, S.C., Kendall, K., High Temperature Solid Oxide Fuel Cells:
Fundamentals, Design, and Applications, Elsevier Advanced Technology,
Oxford, (2003).
294. Fergus, J.W., J. Power Sources, 162 (2006) 30–40.
295. Brett, D.J.L., Atkinson, A., Brandon, N.P., Skinner, S.J., Chem. Soc. Rev.,
37 (2008) 1568–1578.
296. Jacobson, A.J., Chem. Mater., 22 (2010) 660–674.
297. Stiller, C., Thorud, B., Seljebo, S., Mathisen, O., Karoliussen, H.,
Bolland, O., J. Power Sources, 141 (2005) 227–240.
298. Singh, P., Minh, N.Q., Int. J. Appl. Ceram. Tec., 1 (2004) 5–15.
299. Singman, D., J. Electrochem. Soc., 113 (1966) 502–504.
300. Godickemeier, M., Gauckler, L.J., J. Electrochem. Soc., 145 (1998)
414–421.
301. Steele, B., Solid State Ionics, 134 (2000) 3–20.
302. Zhu, B., Yang, X., Xu, J., Zhu, Z., Ji, S., Sun, M., Sun, J., J. Power Sources,
118 (2003) 47–53.
303. Zhu, B., Liu, X., Zhu, Z., Ljungberg, R., Int. J. Hydrogen Energ., 33
(2008) 3385–3392.
304. Gao, Z., Huang, J., Mao, Z., Wang, C., Liu, Z., Int. J. Hydrogen Energ.,
35 (2010) 731–737.
305. Ma, Y., Wang, X., Li, S., Toprak, M.S., Zhu, B., Muhammed, M., Adv.
Mater., 22 (2010) 1640–1644.

b1469_Ch-12.indd 769 4/8/2013 12:41:05 PM


b1469 Catalysis by Ceria and Related Materials

770 C. Chatzichristodoulou et al.

306. Gardner, F., Day, M., Brandon, N., Pashley, M., Cassidy, M., J. Power
Sources, 86 (2000) 122–129.
307. George, R., J. Power Sources, 86 (2000) 134–139.
308. Agnew, G.D., Collins, R.D., Jorger, M., Pyke, S.H., Travis, R.P., SOFC-X,
7 (2007) 105–111.
309. Nagel, F.P., Schildhauer, T.J., Biollaz, S.M.A., Wokaun, A., J. Power
Sources, 184 (2008) 143–164.
310. Evans, A., Bieberle-Huetter, A., Rupp, J.L.M., Gauckler, L.J., J. Power
Sources, 194 (2009) 119–129.
311. Tucker, M.C., J. Power Sources, 195 (2010) 4570–4582.
312. Hibino, T., Hashimoto, A., Inoue, T., Tokuno, J., Yoshida, S., Sano, M.,
Science, 288 (2000) 2031–2033.
313. Shao, Z., Mederos, J., Chueh, W.C., Haile, S.M., J. Power Sources, 162
(2006) 589–596.
314. Kuhn, M., Napporn, T.W., Energies, 3 (2010) 57–134.
315. Takahashi, T., Ito, K., Iwahara, H., Electrochim. Acta, 12 (1967)
21–30.
316. Choudhury, N.S., Patterson, J.W., J. Electrochem. Soc., 118 (1971)
1398–1403.
317. Riess, I., J. Electrochem. Soc., 128 (1981) 2077–2081.
318. Riess, I., J. Phys. Chem. Sol., 47 (1986) 129–138.
319. Yuan, S., Pal, U., J. Electrochem. Soc., 143 (1996) 3214–3222.
320. Näfe, H., J. Appl. Electrochem., 31 (2001) 1235–1241.
321. Yoon, K.J., Huang, W., Ye, G., Gopalan, S., Pal, U.B., Seccombe, D.A.,
J. Electrochem. Soc., 154 (2007) B389–395.
322. Riess, I., Godickemeier, M., Gauckler, L., Solid State Ionics, 90 (1996)
91–104.
323. Godickemeier, M., Sasaki, K., Gauckler, L.J., Riess, I., J. Electrochem.
Soc., 144 (1997) 1635–1646.
324. Soral, P., Pal, U., Worrell, W., J. Electrochem. Soc., 145 (1998) 99–106.
325. Dalslet, B., Blennow, P., Hendriksen, P., Bonanos, N., Lybye, D.,
Mogensen, M., J. Solid State Electr., 10 (2006) 547–561.
326. Serincan, M.F., Pasaogullari, U., Sammes, N.M., J. Electrochem. Soc., 155
(2008) B1117–1127.
327. Duncan, K.L., Lee, K., Wachsman, E.D., J. Power Sources, 196 (2011)
2445–2451.

b1469_Ch-12.indd 770 4/8/2013 12:41:05 PM


b1469 Catalysis by Ceria and Related Materials

Ceria and its Use in Solid Oxide Cells and Oxygen Membranes 771

328. Chatzichristodoulou, C., Sogaard, M., Hendriksen, P.V., J. Electrochem.


Soc., 158 (2011) F61–72.
329. Hao, Y., Goodwin, D.G., J. Electrochem. Soc., 154 (2007) B207–217.
330. Duncan, K.L., Wachsman, E.D., J. Electrochem. Soc., 156 (2009)
B1030–1038.
331. Virkar, A., J. Power Sources, 147 (2005) 8–31.
332. Leah, R.T., Brandon, N.P., Aguiar, P., J. Power Sources, 145 (2005)
336–352.
333. Godickemeier, M., Gauckler, L., J. Electrochem. Soc., 145 (1998)
414–421.
334. Yokokawa, H., Sakai, N., Horita, T., Yamaji, K., Brito, M., MRS Bull., 30
(2005) 591–595.
335. Suzuki, T., Yamaguchi, T., Fujishiro, Y., Awano, M., J. Electrochem. Soc.,
153 (2006) A925–928.
336. Serincan, M.F., Pasaogullari, U., Sammes, N.M., J. Power Sources, 194
(2009) 864–872.
337. Hao, Y., Shao, Z., Mederos, J., Lai, W., Goodwin, D.G., Haile, S.M.,
Solid State Ionics, 177 (2006) 2013–2021.
338. Milliken, C., Guruswamy, S., Khandkar, A., J. Electrochem. Soc., 146
(1999) 872–882.
339. Hatchwell, C., Sammes, N., Brown, I., Solid State Ionics, 126 (1999)
201–208.
340. Suzuki, T., Yamaguchi, T., Fujishiro, Y., Awano, M., J. Power Sources, 160
(2006) 73–77.
341. Suzuki, T., Funahashi, Y., Yamaguchi, T., Fujishiro, Y., Awano, M.,
J. Power Sources, 175 (2008) 68–74.
342. Zhang, X., Robertson, M., Deces-Petit, C., Qu, W., Kesler, O.,
Maric, R., Ghosh, D., J. Power Sources, 164 (2007) 668–677.
343. Chatzichristodoulou, C., Sogaard, M., Glasscock, J., Kaiser, A.,
Foghmoes, S.P.V., Hendriksen, P.V., J. Electrochem. Soc., 158 (2011)
F73–83.
344. Liu, Q.L., Khor, K.A., Chan, S.H., J. Power Sources, 161 (2006) 123–128.
345. Ahn, J.S., Omar, S., Yoon, H., Nino, J.C., Wachsman, E.D., J. Power
Sources, 195 (2010) 2131–2135.
346. Ahn, J.S., Yoon, H., Lee, K.T., Camaratta, M.A., Wachsman, E.D., Fuel
Cells, 9 (2009) 643–649.

b1469_Ch-12.indd 771 4/8/2013 12:41:05 PM


b1469 Catalysis by Ceria and Related Materials

772 C. Chatzichristodoulou et al.

347. Omar, S., Wachsman, E.D., Nino, J.C., Appl. Phys. Lett., 91 (2007) 144106.
348. Ding, C., Lin, H., Sato, K., Amezawa, K., Kawada, T., Mizusaki, J.,
Hashida, T., J. Power Sources, 195 (2010) 5487–5492.
349. Yin, Y., Zhu, W., Xia, C., Gao, C., Meng, G., J. Appl. Electrochem., 34
(2004) 1287–1291.
350. Park, H.C., Virkar, A.V., J. Power Sources, 186 (2009) 133–137.
351. Wang, Z., Berghaus, J.O., Yick, S., Deces-Petit, C., Qu, W., Hui, R.,
Maric, R., Ghosh, D., J. Power Sources, 176 (2008) 90–95.
352. Huang, Q., Wang, B., Qu, W., Hui, R., J. Power Sources, 191 (2009)
297–303.
353. Ai, N., Lu, Z., Chen, K., Huang, X., Liu, Y., Wang, R., Su, W.,
J. Membrane Sci., 286 (2006) 255–259.
354. Berghaus, J.O., Legoux, J., Moreau, C., Hui, R., Deces-Petit, C.,
Qu, W., Yick, S., Wang, Z., Maric, R., Ghosh, D., J. Therm. Spray Techn.,
17 (2008) 700–707.
355. Leng, Y., Chan, S., Jiang, S., Khor, K., Solid State Ionics, 170 (2004)
9–15.
356. Xia, C., Liu, M., Adv. Mater., 14 (2002) 521–523.
357. Liu, Q.L., Chan, S.H., Fu, C.J., Pasciak, G., Electrochem. Commun., 11
(2009) 871–874.
358. Misono, T., Murata, K., Fukui, T., Chaichanawong, J., Sato, K.,
Abe, H., Naito, M., J. Power Sources, 157 (2006) 754–757.
359. Shao, Z., Haile, S., Nature, 431 (2004) 170–173.
360. Hibino, T., Hashimoto, A., Asano, K., Yano, M., Suzuki, M., Sano, M.,
Electrochem. Solid St., 5 (2002) A242–244.
361. Yang, M., Yan, A., Zhang, M., Hou, Z., Dong, Y., Cheng, M., J. Power
Sources, 175 (2008) 345–352.
362. Zha, S., Moore, A., Abernathy, H., Liu, M., J. Electrochem. Soc., 151
(2004) A1128–1133.
363. Brandon, N.P., Corcoran, D., Cummins, D., Duckett, A., El-Khoury, K.,
Haigh, D., Leah, R., Lewis, G., Maynard, N., McColm, T., Trezona, R.,
Selcuk, A., Schmidt, M., J. Mater. Eng. Perform., 13 (2004) 253–256.
364. Bance, P., Brandon, N., Girvan, B., Holbeche, P., O’Dea, S., Steele, B.,
J. Power Sources, 131 (2004) 86–90.
365. Huijsmans, J., Berkel, F.V., Christie, G., J. Power Sources, 71 (1998)
107–110.

b1469_Ch-12.indd 772 4/8/2013 12:41:05 PM


b1469 Catalysis by Ceria and Related Materials

Ceria and its Use in Solid Oxide Cells and Oxygen Membranes 773

366. Zhang, X., Robertson, M., Yick, S., Deces-Petit, C., Styles, E., Qu, W.,
Xie, Y., Hui, R., Roller, J., Kesler, O., Maric, R., Ghosh, D., J. Power
Sources, 160 (2006) 1211–1216.
367. Tuck, A., Zhang, X., Hui, R., Qu, W., Deces-Petit, C., Xie, Y., Roller, J.,
Min, K., Robertson, M., Yick, S., Maric, R., Ghosh, D., SOFC-X, 7
(2007) 285–294.
368. Hui, S., Yang, D., Wang, Z., Yick, S., Deces-Petit, C., Qu, W., Tuck, A.,
Maric, R., Ghosh, D., J. Power Sources, 167 (2007) 336–339.
369. Xie, Y., Neagu, R., Hsu, C., Zhang, X., Deces-Petit, C., Qu, W., Hui, R.,
Yick, S., Robertson, M., Maric, R., Ghosh, D., J. Fuel Cell Sci. Tech., 7
(2010) 021007.
370. Wang, W., Mogensen, M., Solid State Ionics, 176 (2005) 457–462.
371. Mai, A., Haanappel, V., Uhlenbruck, S., Tietz, F., Stover, D., Solid State
Ionics, 176 (2005) 1341–1350.
372. Hjelm, J., Sogaard, M., Wandel, M., Menon, M., Mogensen, M.,
Hagen, A., SOFC-X, 7 (2007) 1261–1270.
373. Simner, S., Shelton, J., Anderson, M., Stevenson, J., Solid State Ionics,
161 (2003) 11–18.
374. Uhlenbruck, S., Moskalewicz, T., Jordan, N., Penkalla, H.,
Buchkremer, H.P., Solid State Ionics, 180 (2009) 418–423.
375. Knibbe, R., Hjelm, J., Menon, M., Pryds, N., Sogaard, M., Wang, H.J.,
Neufeld, K., J. Am. Ceram. Soc., 93 (2010) 2877–2883.
376. Tsoga, A., Gupta, A., Naoumidis, A., Nikolopoulos, P., Acta Mater., 48
(2000) 4709–4714.
377. Zhou, X., Scarfino, B., Anderson, H., Solid State Ionics, 175 (2004)
19–22.
378. Uchida, H., Arisaka, S., Watanabe, M., Electrochem. Solid St., 2 (1999)
428–430.
379. Simner, S., Bonnett, J., Canfield, N., Meinhardt, K., Sprenkle, V.,
Stevenson, J., Electrochem. Solid St., 5 (2002) A173–175.
380. Mai, A., Haanappel, V.A.C., Tietz, F., Stoever, D., Solid State Ionics, 177
(2006) 2103–2107.
381. Uhlenbruck, S., Jordan, N., Sebold, D., Buchkremer, H.P., Haanappel,
V.A.C., Stoever, D., Thin Solid Films, 515 (2007) 4053–4060.
382. Plonczak, P., Joost, M., Hjelm, J., Sogaard, M., Lundberg, M.,
Hendriksen, P.V., J. Power Sources, 196 (2011) 1156–1162.

b1469_Ch-12.indd 773 4/8/2013 12:41:05 PM


b1469 Catalysis by Ceria and Related Materials

774 C. Chatzichristodoulou et al.

383. Endler-Schuck, C., Weber, A., Ivers-Tiffee, E., Guntow, U., Ernst, J.,
Ruska, J., J. Fuel Cell Sci. Tech., 8 (2011) 041001.
384. Fonseca, F.C., Uhlenbruck, S., Nedelec, R., Sebold, D., Buchkremer, H.P.,
J. Electrochem. Soc., 157 (2010) B1515–1519.
385. Nguyen, T., Kobayashi, K., Honda, T., Iimura, Y., Kato, K., Neghisi, A.,
Nozaki, K., Tappero, F., Sasaki, K., Shirahama, H., Ota, K., Dokiya, M.,
Kato, T., Solid State Ionics, 174 (2004) 163–174.
386. Kim, W., Song, H., Moon, J., Lee, H., Solid State Ionics, 177 (2006)
3211–3216.
387. Wang, Z., Hashimoto, S., Mori, M., J. Power Sources, 193 (2009) 49–54.
388. Kungas, R., Bidrawn, F., Vohs, J.M., Gorte, R.J., Electrochem. Solid St., 13
(2010) B87–90.
389. Song, J., Park, Y., Bae, H., Ahn, J., Seong, B., Kim, D., Jun, J., in
Proceedings of the ASME 8th International Conference on Fuel Cell Science,
Engineering, and Technology, American Society of Mechanical Engineers,
Brooklyn, NY, (2010), p. 319.
390. Jordan, N., Assenmacher, W., Uhlenbruck, S., Haanappel, V.A.C.,
Buchkremer, H.P., Stoever, D., Mader, W., Solid State Ionics, 179 (2008)
919–923.
391. Emsley, J. Nature’s Building Blocks: An A–Z Guide to the Elements, Oxford
University Press, New York, (2001).
392. Bennaceur, K., Gielen, D., Kerr, T., Tam, C., CO2. Capture and Storage:
A Key Carbon Abatement Option, IEA Publications, Paris, (2008).
393. Sunarso, J., Baumann, S., Serra, J.M., Meulenberg, W.A., Liu, S.,
Lin, Y.S., da Costa, J.C.D., J. Membrane Sci., 320 (2008) 13–41.
394. Armstrong, P.A., Bennet, D.L., Foster, E.P., Stein, V.E.E., ITM Oxygen:
The New Oxygen Supply for the New IGCC Market, Paper presented at
Gasification Technologies 2005 Conference: New Technology
Developments, San Francisco, CA, (2005).
395. http://www.praxair.com/praxair.nsf/0/179a71da389ab4bb852577a7
004d7120?OpenDocument. Accessed 20 December 2012.
396. Gellings, P.J., Bouwmeester, H.J.M., in The CRC Handbook of Solid State
Electrochemistry, eds. H.J.M. Bouwmeester, Gellings, CRC Press, Boca
Raton, (1997).
397. Hendriksen, P., Larsen, P., Mogensen, M., Poulsen, F., Wiik, K., Catal.
Today, 56 (2000) 283–295.

b1469_Ch-12.indd 774 4/8/2013 12:41:05 PM


b1469 Catalysis by Ceria and Related Materials

Ceria and its Use in Solid Oxide Cells and Oxygen Membranes 775

398. Sogaard, M., Hendriksen, P.V., Mogensen, M., J. Solid State Chem., 180
(2007) 1489–1503.
399. Engels, S., Beggel, F., Modigell, M., Stadler, H., J. Membrane Sci., 359
(2010) 93–101.
400. Engels, S., Modigell, M., Desalination, 199 (2006) 291–292.
401. http://www.chemmarket.info/en/home/article/1872/. Accessed 20
December 2012.
402. http://methanol.org/getattachment/Home/Methanol-Fuel-and-
Flexible-Fuel-Technology.pdf.aspx. Accessed 20 December 2012.
403. Park, H., Choi, G., J. Eur. Ceram. Soc., 24 (2004) 1313–1317.
404. Akin, F., Lin, Y., Catal. Lett., 78 (2002) 239–242.
405. Teraoka, Y., Zhang, H., Furukawa, S., Yamazoe, N., Chem. Lett., 11
(1985) 1743–1746.
406. Tsai, C., Dixon, A.G., Ma, Y.H., Moser, W.R., Pascucci, M.R., J. Am.
Ceram. Soc., 81 (1998) 1437–1444.
407. Wang, H., Cong, Y., Yang, W., J. Membrane Sci., 210 (2002) 259–271.
408. Lein, H.L., Wiik, K., Einarsrud, M., Grande, T., Lara-Curzio, E., J. Am.
Ceram. Soc., 89 (2006) 2895–2898.
409. Lein, H.L., Wiik, K., Grande, T., Solid State Ionics, 177 (2006)
1587–1590.
410. Sogaard, M., Hendriksen, P.V., Mogensen, M., Poulsen, F.W., Skou, E.,
Solid State Ionics, 177 (2006) 3285–3296.
411. Baumann, S., Serra, J.M., Lobera, M.P., Escolastico, S., Schulze-
Kueppers, F., Meulenberg, W.A., J. Membrane Sci., 377 (2011)
198–205.
412. Hendriksen, P.V., Hogsberg, J.R., Kjeldsen, A.M., Sorensen, B.F.,
Pedersen, H.G., Ceram. Eng. Sci. Proc., 27 (2006) 347.
413. Niedrig, C., Taufall, S., Burriel, M., Menesklou, W., Wagner, S.F.,
Baumann, S., Ivers-Tiffee, E., Solid State Ionics, 197 (2011) 25–31.
414. Kovalevsky, A.V., Yaremchenko, A.A., Kolotygin, V.A., Snijkers, F.M.M.,
Kharton, V.V., Buekenhoudt, A., Luyten, J.J., Solid State Ionics, 192
(2011) 677–681.
415. Jiang, Q., Faraji, S., Nordheden, K.J., Stagg-Williams, S.M., J. Membrane
Sci., 368 (2011) 69–77.
416. Meixner, D., Brengel, D., Henderson, B., Abrardo, J., Wilson, M.,
Taylor, D., Cutler, R., J. Electrochem. Soc., 149 (2002) D132–136.

b1469_Ch-12.indd 775 4/8/2013 12:41:05 PM


b1469 Catalysis by Ceria and Related Materials

776 C. Chatzichristodoulou et al.

417. Hutchings, K.N., Bai, J., Cutler, R.A., Wilson, M.A., Taylor, D.M., Solid
State Ionics, 179 (2008) 442–450.
418. Roettger, B.F., in Proceedings of the Annual Survival and Flight Equipment
Association Symposium, 29 (1991) 247.
419. Park, H.J., Choi, G.M., J. Eur. Ceram. Soc., 24 (2004) 1313–1317.
420. Park, H.J., Choi, G.M., Solid State Ionics, 178 (2008) 1746–1755.
421. Yin, X., Hong, L., Liu, Z., Appl. Catal. A: Gen., 300 (2006) 75–84.
422. Kaiser, A., Foghmoes, S., Chatzichristodoulou, C., Sogaard, M.,
Glasscock, J.A., Frandsen, H.L., Hendriksen, P.V., J. Membrane Sci., 378
(2011) 51–60.
423. Atkinson, A., Solid State Ionics, 95 (1997) 249–258.
424. Takamura, H., Kobayashi, T., Kasahara, T., Kamegawa, A., Okada, M.,
J. Alloy Compd., 408 (2006) 1084–1089.
425. Zhu, X., Yang, W., AIChE J., 54 (2008) 665–672.
426. Kharton, V., Kovalevsky, A., Viskup, A., Shaula, A., Figueiredo, F.,
Naumovich, E., Marques, F., Solid State Ionics, 160 (2003) 247–258.
427. Schmale, K., PhD Thesis, University of Münster (2011).
428. Perez-Coll, D., Nunez, P., Abrantes, J.C.C., Fagg, D.P., Kharton, V.V.,
Frade, J.R., Solid State Ionics, 176 (2005) 2799–2805.
429. Fagg, D.P., Shaula, A.L., Kharton, V.V., Frade, J.R., J. Membrane Sci.,
299 (2007) 1–7.
430. Kharton, V., Kovalevsky, A., Viskup, A., Figueiredo, F., Yaremchenko, A.,
Naumovich, E., Marques, F., J. Electrochem. Soc., 147 (2000) 2814–2821.
431. Kobayashi, K., Nishioka, M., Sato, K., Inoue, T., Hamakawa, S.,
Tsunoda, T., J. Solid State Electr., 10 (2006) 629–634.
432. Yi, J., Zuo, Y., Liu, W., Winnubst, L., Chen, C., J. Membrane Sci., 280
(2006) 849–855.
433. Zhu, X., Li, Q., He, Y., Cong, Y., Yang, W., J. Membrane Sci., 360 (2010)
454–460.
434. Chen, T., Zhao, H., Xu, N., Li, Y., Lu, X., Ding, W., Li, F., J. Membrane
Sci., 370 (2011) 158–165.
435. Kagomiya, I., Iijima, T., Takamura, H., J. Membrane Sci., 286 (2006)
180–184.
436. Wang, B., Yi, J., Winnubst, L., Chen, C., J. Membrane Sci., 286 (2006)
22–25.

b1469_Ch-12.indd 776 4/8/2013 12:41:05 PM


b1469 Catalysis by Ceria and Related Materials

Ceria and its Use in Solid Oxide Cells and Oxygen Membranes 777

437. Li, W., Tian, T., Shi, F., Wang, Y., Chen, C., Ind. Eng. Chem. Res., 48
(2009) 5789–5793.
438. Yin, X., Hong, L., Liu, Z., J. Membrane Sci., 268 (2006) 2–12.
439. Mazanec, T., Cable, T., Frye, J., Solid State Ionics, 53 (1992) 111–118.
440. Ji, Y., Kilner, J., Carolan, M., J. Eur. Ceram. Soc., 24 (2004) 3613–3616.
441. Frolova-Borchert, Y.V., Sadykov, V.A., Alikina, G.M., Lukashevich, A.I.,
Moroz, E.M., Kochubey, D.I., Kriventsov, V.V., Zaikovskii, V.I.,
Zyryanov, V.V., Uvarov, N.F., Solid State Ionics, 177 (2006) 2533–2538.
442. Seeharaj, P., Berenov, A., Raj, E., Rudkin, R., Atkinson, A., Solid State
Ionics, 192 (2011) 638–641.
443. Forskningscenter Riso, Danish Patent, WO2002041434A1, 05/23.
444. Takamura, H., Ogawa, M., Suehiro, K., Takahashi, H., Okada, M.,
Solid State Ionics, 179 (2008) 1354–1359.
445. Kharton, V.V., Figueiredo, F.M., Navarro, L., Naumovich, E.N.,
Kovalevsky, A.V., Yaremchenko, A.A., Viskup, A.P., Carneiro, A.,
Marques, F.M.B., Frade, J.R., J. Mater. Sci., 36 (2001) 1105–1117.
446. Jud, E., Zhang, Z., Sigle, W., Gauckler, L., J. Electroceram., 16 (2006)
191–197.
447. Takamura, H., Okumura, K., Koshino, Y., Kamegawa, A., Okada, M.,
J. Electroceram., 13 (2004) 613–618.
448. Mogensen, M., Kammer, K., Ann. Rev. Mater. Res., 33 (2003)
321–331.
449. Steghuis, A.G., van Ommen, J.G., Lercher, J.A., Catal. Today, 46 (1998)
91–97.
450. Doornkamp, C., Ponec, V., J. Mol. Catal. A: Chem., 162 (2000) 19–32.
451. Adler, S.B., Chem. Rev., 104 (2004) 4791–4843.
452. Sun, C., Hui, R., Roller, J., J. Solid State Electr., 14 (2010) 1125–1144.
453. Matsuzaki, Y., Yasuda, I., Solid State Ionics, 132 (2000) 261–269.
454. Xiong, Y., Yamaji, K., Horita, T., Yokokawa, H., Akikusa, J., Eto, H.,
Inagaki, T., J. Electrochem. Soc., 156 (2009) B588–592.
455. Pihlatie, M., Kaiser, A., Larsen, P.H., Mogensen, M., J. Electrochem. Soc.,
156 (2009) B322–329.
456. Pihlatie, M., Kaiser, A., Mogensen, M., Solid State Ionics, 180 (2009)
1100–1112.
457. Tsipis, E.V., Kharton, V.V., J. Solid State Electr., 12 (2008) 1367–1391.

b1469_Ch-12.indd 777 4/8/2013 12:41:05 PM


b1469 Catalysis by Ceria and Related Materials

778 C. Chatzichristodoulou et al.

458. Steele, B.C.H., Kelly, I., Middleton, H., Rudkin, R., Solid State Ionics,
28–30 (1988) 1547–1552.
459. Gorte, R.J., Vohs, J.M., J. Catal., 216 (2003) 477–486.
460. Mitterdorfer, A., Gauckler, L.J., Solid State Ionics, 111 (1998) 185–218.
461. Liu, Y.L., Hagen, A., Barfod, R., Chen, M., Wang, H.J., Poulsen, F.W.,
Hendriksen, P.V., Solid State Ionics, 180 (2009) 1298–1304.
462. Chatzichristodoulou, C., Chen, M., Bowen, J.R., Liu, Y., J. Am. Ceram.
Soc., 93 (2010) 2884–2890.
463. Primdahl, S., Mogensen, M., Solid State Ionics, 152–153 (2002)
597–608.
464. Mogensen, M., Primdahl, S., Jørgensen, M.J., Bagger, C., J. Electroceram.,
5 (2000) 141–152.
465. Tsipis, E.V., Kharton, V.V., J. Solid State Electr., 12 (2008) 1039–1060.
466. Takahashi, T., Iwahara, H., Suzuki, Y., in Proceedings of the Third
International Symposium on Fuel Cells, Presses Académiques
Européennes, Brussels, (1969) p. 114.
467. Möbius, H.H. and Rohland, B. US Patent No. 3,377,203, April 9, 1968.
468. Mogensen, M., Bentzen, J.J., in Proceedings of the 1st International
Symposium on Solid Oxide Fuel Cells, 89 (1989) 99.
469. Steele, B.C.H., Middleton, P.H., Rudkin, R.A., Solid State Ionics, 40–41
(1990) 388–393.
470. Putna, E.S., Stubenrauch, J., Vohs, J.M., Gorte, R.J., Langmuir, 11
(1995) 4832–4837.
471. Marina, O.A., Bagger, C., Primdahl, S., Mogensen, M., Solid State
Ionics, 123 (1999) 199–208.
472. Uchida, H., Suzuki, H., Watanabe, M., J. Electrochem. Soc., 145 (1998)
615–620.
473. Primdahl, S., Liu, Y.L., J. Electrochem. Soc., 149 (2002) A1466–1472.
474. Hammer, B., Norskov, J., Surf. Sci., 343 (1995) 211–220.
475. Gross, A., Surf. Sci. Rep., 32 (1998) 291–340.
476. Greeley, J., Mavrikakis, M., Nat. Mater., 3 (2004) 810–815.
477. Chueh, W.C., Hao, Y., Jung, W., Haile, S.M., Nat. Mater., 11 (2011),
155–161.
478. Lai, W., Haile, S.M., J. Am. Ceram. Soc., 88 (2005) 2979–2997.
479. Chueh, W.C., Lai, W., Haile, S.M., Solid State Ionics, 179 (2008)
1036–1041.

b1469_Ch-12.indd 778 4/8/2013 12:41:05 PM


b1469 Catalysis by Ceria and Related Materials

Ceria and its Use in Solid Oxide Cells and Oxygen Membranes 779

480. Zhang, C., Grass, M.E., McDaniel, A.H., Decaluwe, S.C., Gabaly, F.E.,
Liu, Z., McCarty, K.F., Farrow, R.L., Linne, M.A., Hussain, Z., Jackson, G.S.,
Bluhm, H., Eichhorn, B.W., Nat. Mater., 9 (2010) 944–949.
481. Schouler, E.J.L., Kleitz, M., J. Electrochem. Soc., 134 (1987) 1045–1050.
482. Uchida, H., Mochizuki, N., Watanabe, M., J. Electrochem. Soc., 143
(1996) 1700–1704.
483. Watanabe, M., Uchida, H., Yoshida, M., J. Electrochem. Soc., 144 (1997)
1739–1743.
484. Tsai, T., Barnett, S.A., J. Electrochem. Soc., 145 (1998) 1696–1701.
485. Horita, T., Yamaji, K., Sakai, N., Ishikawa, M., Yokokawa, H., Kawada, T.,
Dokiya, M., Electrochem. Solid St., 1 (1998) 4–6.
486. Bentzen, J.J., Schwartzbach, H., Solid State Ionics, 40–41 (1990)
942–946.
487. Perry Murray, E., Tsai, T., Barnett, S.A., Nature, 400 (1999) 649–651.
488. Park, S., Vohs, J.M., Gorte, R.J., Nature, 404 (2000) 265–267.
489. Park, S., Gorte, R.J., Vohs, J.M., Appl. Catal. A: Gen., 200 (2000) 55–61.
490. Gorte, R.J., Park, S., Vohs, J.M., Wang, C., Adv. Mater., 12 (2000)
1465–1469.
491. Gorte, R.J., Kim, H., Vohs, J.M., J. Power Sources, 106 (2002) 10–15.
492. Liu, J., Madsen, B.D., Ji, Z., Barnett, S.A., Electrochem. Solid St., 5 (2002)
A122–124.
493. Marina, O.A., Mogensen, M., Appl. Catal. A: Gen., 189 (1999) 117–126.
494. McIntosh, S., Gorte, R.J., Chem. Rev., 104 (2004) 4845–4865.
495. Aguiar, P., Ramírez-Cabrera, E., Lapeña-Rey, N., Atkinson, A.,
Kershenbaum, L.S., Chadwick, D., in Studies in Surface Science and
Catalysis 136: Natural Gas Conversion VI, eds. E. Iglesia, J.J. Spivey, T.H.
Fleisch, Elsevier, Amsterdam, (2001), p. 501.
496. Aguiar, P., Ramírez-Cabrera, E., Atkinson, A., Kershenbaum, L.S.,
Chadwick, D., in Proceedings of the 7th International Symposium on Solid
Oxide Fuel Cells, 16 (2001) 703.
497. Ramírez-Cabrera, E., Atkinson, A., Chadwick, D., Appl. Catal. B:
Environ., 47 (2004) 127–131.
498. Kammer, K., Mogensen, M., Electrochem. Solid St., 8 (2005) A108–109.
499. González-Velasco, J.R., Gutiérrez-Ortiz, M.A., Marc, J., Botas, J.A.,
González-Marcos, M.P., Blanchard, G., Appl. Catal. B: Environ., 22
(1999) 167–178.

b1469_Ch-12.indd 779 4/8/2013 12:41:05 PM


b1469 Catalysis by Ceria and Related Materials

780 C. Chatzichristodoulou et al.

500. Jen, H., Graham, G.W., Chun, W., McCabe, R.W., Cuif, J., Deutsch,
S.E., Touret, O., Catal. Today, 50 (1999) 309–328.
501. Hori, C.E., Permana, H., Ng, K.Y.S., Brenner, A., More, K., Rahmoeller,
K.M., Belton, D., Appl. Catal. B: Environ., 16 (1998) 105–117.
502. O’Connell, M., Morris, M.A., Catal. Today, 59 (2000) 387–393.
503. Ramírez-Cabrera, E., Atkinson, A., Chadwick, D., Appl. Catal. B:
Environ., 36 (2002) 193–206.
504. Zhao, S., Gorte, R.J., Appl. Catal. A: Gen., 277 (2004) 129–136.
505. McIntosh, S., Vohs, J.M., Gorte, R.J., Electrochim. Acta, 47 (2002)
3815–3821.
506. Wang, J.B., Jang, J., Huang, T., J. Power Sources, 122 (2003) 122–131.
507. Tao, S., Irvine, J.T.S., Chem. Rec., 4 (2004) 83–95.
508. Atkinson, A., Barnett, S., Gorte, R.J., Irvine, J.T.S., McEvoy, A.J.,
Mogensen, M., Singhal, S.C., Vohs, J., Nat. Mater., 3 (2004) 17–27.
509. Jiang, S.P., Mat. Sci. Eng. A: Struct., 418 (2006) 199–210.
510. Sholklapper, T.Z., Kurokawa, H., Jacobson, C.P., Visco, S.J., De
Jonghe, L.C., Nano Lett., 7 (2007) 2136–2141.
511. Jiang, Z., Xia, C., Chen, F., Electrochim. Acta, 55 (2010) 3595–3605.
512. Gorte, R.J., Vohs, J.M., Ann.Rev.: Chem. Biomol. Eng., 2 (2011) 9–30.
513. Hojberg, J., Sogaard, M., Electrochem. Solid St., 14 (2011) B77–79.
514. Knöfel, C., Wang, H., Thydén, K.T.S., Mogensen, M., Solid State Ionics,
195 (2011) 36–42.
515. Mogensen, M., Sogaard, M., Blennow, P., Hansen, K.K., in Proceedings
of the 8th European Solid Oxide Fuel Cell Forum, eds. R. Steinberger-
Wilckens and U. Bossel, Lucerne Fuel Cell Forum, Switzerland,
(2008) A0402.
516. Kurokawa, H., Sholklapper, T.Z., Jacobson, C.P., De Jonghe, L.C.,
Visco, S.J., Electrochem. Solid St., 10 (2007) B135–138.
517. Yun, J.W., Yoon, S.P., Park, S., Kim, H.S., Nam, S.W., Int. J. Hydrogen
Energ., 36 (2011) 787–796.
518. Ferrizz, R.M., Gorte, R.J., Vohs, J.M., Appl. Catal. B: Environ., 43 (2003)
273–280.
519. Lohsoontorn, P., Brett, D.J.L., Brandon, N.P., J. Power Sources, 175
(2008) 60–67.
520. Zeng, Y., Kaytakoglu, S., Harrison, D.P., Chem. Eng. Sci., 55 (2000)
4893–4900.

b1469_Ch-12.indd 780 4/8/2013 12:41:05 PM


b1469 Catalysis by Ceria and Related Materials

Ceria and its Use in Solid Oxide Cells and Oxygen Membranes 781

521. Kim, H., Vohs, J.M., Gorte, R.J., Chem. Commun., 22 (2001) 2334–2335.
522. Gorte, R.J., Vohs, J.M., McIntosh, S., Solid State Ionics, 175 (2004) 1–6.
523. Brett, D.J.L., Atkinson, A., Cumming, D., Ramírez-Cabrera, E.,
Rudkin, R., Brandon, N.P., Chem. Eng. Sci., 60 (2005) 5649–5662.
524. Ye, X., Wang, S.R., Zhou, J., Zeng, F.R., Nie, H.W., Wen, T.L., J. Power
Sources, 195 (2010) 7264–7267.
525. Haryanto, A., Fernando, S., Murali, N., Adhikari, S., Energy and Fuels,
19 (2005) 2098–2106.
526. Lee, S., Vohs, J.M., Gorte, R.J., J. Electrochem. Soc., 151 (2004)
A1319–1323.
527. Lee, S., Ahn, K., Vohs, J.M., Gorte, R.J., Electrochem. Solid St., 8 (2005)
A48–51.
528. Gross, M.D., Vohs, J.M., Gorte, R.J., Electrochim. Acta, 52 (2007)
1951–1957.
529. Tao, S., Irvine, J.T.S., Nat. Mater., 2 (2003) 320–323.
530. Tao, S., Irvine, J.T.S., J. Electrochem. Soc., 151 (2004) A252–259.
531. Jiang, S.P., Chen, X.J., Chan, S.H., Kwok, J.T., J. Electrochem. Soc., 153
(2006) A850–856.
532. Chen, X.J., Liu, Q.L., Chan, S.H., Brandon, N.P., Khor, K.A.,
J. Electrochem. Soc., 154 (2007) B1206–1210.
533. Chen, X.J., Liu, Q.L., Chan, S.H., Brandon, N.P., Khor, K.A.,
Electrochem. Commun., 9 (2007) 767–772.
534. Kim, J., Nair, V.V., Vohs, J.M., Gorte, R.J., Scripta Mater., 65 (2011)
90–95.
535. Jiang, S.P., Chen, X.J., Chan, S.H., Kwok, J.T., Khor, K.A., Solid State
Ionics, 177 (2006) 149–157.
536. Blennow, P., Hagen, A., Hansen, K.K., Wallenberg, L.R., Mogensen, M.,
Solid State Ionics, 179 (2008) 2047–2058.
537. Blennow, P., Hansen, K.K., Wallenberg, L.R., Mogensen, M., ECS
Transactions, 13 (2008) 181–194.
538. Gross, M.D., Carver, K.M., Deighan, M.A., Schenkel, A., Smith, B.M.,
Yee, A.Z., J. Electrochem. Soc., 156 (2009) B540–545.
539. Kurokawa, H., Yang, L., Jacobson, C.P., De Jonghe, L.C., Visco, S.J.,
J. Power Sources, 164 (2007) 510–518.
540. Savaniu, C.D., Irvine, J.T.S., Solid State Ionics, 192 (2011) 491–493.

b1469_Ch-12.indd 781 4/8/2013 12:41:05 PM


b1469 Catalysis by Ceria and Related Materials

782 C. Chatzichristodoulou et al.

541. Blennow, P., Hjelm, J., Klemensø, T., Persson, Å.H., Ramousse, S.,
Mogensen, M., Fuel Cells, 11 (2011) 661–668.
542. Klemensø, T., Nielsen, J., Blennow, P., Persson, Å.H., Stegk, T.,
Christensen, B.H., Sønderby, S., J. Power Sources, 196 (2011) 9459–9466.
543. Blennow, P., Hjelm, J., Klemensø, T., Ramousse, S., Kromp, A.,
Leonide, A., Weber, A., J. Power Sources, 196 (2011) 7117–7125.
544. Blennow, P., Klemenso, T., Persson, A., Brodersen, K., Srivastava, A.K.,
Sudireddy, B.R., Ramousse, S., Mogensen, M., ECS Transactions, 35
(2011) 683–692.
545. Lane, J.A., Kilner, J.A., Solid State Ionics, 136–137 (2000) 927–932.
546. Armstrong, E.N., Duncan, K.L., Oh, D.J., Weaver, J.F., Wachsman,
E.D., J. Electrochem. Soc., 158 (2011) B492–499.
547. Karthikeyan, A., Ramanathan, S., Appl. Phys. Lett., 92 (2008) 243109.
548. Wang, L., Zhang, C., Grass, M., Yu, Y., Gaskell, K., Hussain, Z., Liu, Z.,
Eichhorn, B., Jackson, G., ECS Transactions, 35 (2011) 1435–1444.
549. DeCaluwe, S.C., Jackson, G., ECS Transactions, 35 (2011) 2883–2895.
550. Chen, H., Chen, Y., Aleksandrov, A., Dong, J., Liu, M., Orlando, T.M.,
Appl. Surf. Sci., 243 (2005) 166–177.
551. Green, R.D., Liu, C., Adler, S.B., Solid State Ionics, 179 (2008)
647–660.
552. Nakamura, T., Kobayashi, T., Yashiro, K., Kaimai, A., Otake, T., Sato, K.,
Mizusaki, J., Kawada, T., J. Electrochem. Soc., 155 (2008) B563–569.
553. Yokokawa, H., Horita, T., Sakai, N., Yamaji, K., Brito, M.E., Xiong, Y.,
Kishimoto, H., Solid State Ionics, 174 (2004) 205–221.
554. Tang, L., Salamon, M., de Guire, M.R., Sci. Adv. Mat., 2 (2010) 79–89.
555. He, H., Vohs, J.M., Gorte, R.J., J. Electrochem. Soc., 150 (2003)
A1470–1475.
556. Watkins, M.B., Foster, A.S., Shluger, A.L., J. Phys. Chem. C, 111 (2007)
15337–15341.
557. Knapp, D., Ziegler, T., J. Phys. Chem. C, 112 (2008) 17311–17318.
558. Skårman, B., Grandjean, D., Benfield, R.E., Hinz, A., Andersson, A.,
Wallenberg, L.R., J. Catal., 211 (2002) 119–133.
559. Lundberg, M., Skårman, B., Wallenberg, L.R., Micropor. Mesopor. Mat.,
69 (2004) 187–195.
560. Aneggi, E., Llorca, J., Boaro, M., Trovarelli, A., J. Catal., 234 (2005)
88–95.

b1469_Ch-12.indd 782 4/8/2013 12:41:05 PM


b1469 Catalysis by Ceria and Related Materials

CHAPTER 13

TRANSFORMATION OF OXYGENATED
COMPOUNDS DERIVED FROM BIOMASS
INTO VALUABLE CHEMICALS USING
CERIA-BASED SOLID CATALYSTS
Laurence Vivier and Daniel Duprez
LACCO, Laboratoire de Catalyse en Chimie Organique
Université de Poitiers & CNRS
4, Rue Michel Brunet, 86022 POITIERS Cedex, France

13.1 Introduction
Ceria is an oxide widely employed in catalysis for its redox proper-
ties.1,2 It also has weak Lewis acid sites and stronger basic sites, which
makes this oxide suitable as a catalyst or a support for many reac-
tions requiring acid–base pairs or basic sites.3 Combining redox and
acid–base properties allows ceria-containing solids to catalyze
numerous reactions involved in biomass transformation. The objec-
tive of this chapter is to review these reactions stressing the role of
ceria-based solid catalysts in the conversion of oxygenated mole-
cules, which are essential compounds in biomass.

13.2 Dehydrogenation of Ethanol to Acetaldehyde


Bioethanol, an environmentally friendly renewable liquid biofuel,
can be produced from several different types of biomass feedstock

783

b1469_Ch-13.indd 783 4/8/2013 12:41:50 PM


b1469 Catalysis by Ceria and Related Materials

784 L. Vivier and D. Duprez

such as sugar or starch from sugar cane, sugar beet, corn and wheat,
and from lignocellulosic biomass. Bioethanol can be produced from
cellulosic residues, which are found in large quantities all over the
world. The use of ethanol as a feedstock for the chemical industry
can be foreseen.
Idriss and collaborators4–7 studied the reactions of ethanol over
noble-metal-based catalysts (Pd, Pt, Rh, and Au) supported on CeO2
by temperature-programmed desorption (TPD), infra-red (IR)
adsorption, and steady-state reactions. The surface areas of the cata-
lysts are given in Table 13.1. In the presence of Pd, Pt, or Au the
dehydrogenation reaction yields acetaldehyde from ethanol.
Acetaldehyde undergoes β-aldolization to form adsorbed crotonal-
dehyde species. Besides acetaldehyde, acetone and benzene were
also observed by TPD. Acetone formation on Pd/CeO2 may be
explained by acetate ketonization on the oxide support or by an
acetyl reaction with methyl species in the presence of Pd or Pt.
A reaction between adsorbed crotonaldehyde and acetaldehyde (via
the same β-aldolization) can explain the formation of benzene in
the vicinity of metal atoms.4,5 Raskó and Kiss showed that these reac-
tions also occurred on oxide support such as CeO2 or TiO2.8 Unlike
Pd/CeO2 and Pt/CeO2, Rh/CeO2 readily dissociates the C–C bond
of ethanol to produce adsorbed CO and CH4.6
More recently Nair et al.9 studied the activity of vanadium oxide
catalysts supported on Al2O3, TiO2, and CeO2 in the ethanol to acet-
aldehyde reaction in the oxidative dehydrogenation (ODH) process
at 180°C. The acetaldehyde formation rate and product selectivity

Table 13.1 BET surface area determination and turnover number (TON) values.
Catalyst BET surface area (m2·g−1) TON (s−1)a Reference
1 wt% Pd/CeO2 55 8.6 4
1 wt% Pt/CeO2 63 2.6 5
1 wt% Rh/CeO2 49 5.9 6
1 wt% Au/CeO2 29 — 7
a
Number of ethanol molecules converted per surface metal atom (Pd, Pt, or Rh) per second
at 127 °C.

b1469_Ch-13.indd 784 4/8/2013 12:41:51 PM


b1469 Catalysis by Ceria and Related Materials

Transformation of Oxygenated Compounds 785

Table 13.2 Acetaldehyde (AC) rates from ethanol per metal


atom and acetaldehyde selectivity for vanadium oxide catalysts
on different supports at 180 °C.9
AC rate × 103
Catalyst a
(mol AC.(metal atoms)−1) AC selectivity (%)
VOx/Al2O3 0.60 67
VOx/CeO2 5.83 100
VOx/TiO2 21.70 100
a 2.
2 vanadium atoms/nm

depend significantly on the support. CeO2-supported and TiO2-


supported VOx are more active and selective than Al2O3-supported
VOx (Table 13.2.). The electronegativity of the support atoms is
involved in the availability and density of active oxygen atoms in the
lattice for acetaldehyde formation; the TiO2 and CeO2 supports are
reducible during the ODH reaction.
Acetaldehyde can also be generated from ethanol on Cu/ZnO/
Al2O3 (CZA) at 150°C.10 The selectivity to ethyl acetate increases in
experiments where the catalyst was in a physical mixture with various
oxides such as ZrO2 (110 m2·g−1), CeO2 (86 m2·g−1), Al2O3 (294 m2·g−1),
and SiO2 (261 m2·g−1). There is a synergy between CZA and the
oxides, which is more intense for ZrO2 and CeO2. These oxides show
strong basic sites (with a density of about 35 µmol·g−1), which gener-
ate the most active and selective systems for ethyl acetate synthesis.10
Catalysts containing redox-active metal oxides of V, Mo, Cu, and Ce,
among others, are useful for the commercial production of acetal-
dehyde from ethanol via ODH.
Bioethanol (20 vol% in aqueous solution) has also been
converted into hydrocarbons over Ce-modified ZSM-5 zeolites.11
The reaction was carried out in a fixed-bed reactor under atmospheric
pressure, at 400°C. Various reactions, such as dehydration, oligomer-
ization, cracking, hydrogen transfer, dehydrocyclization, aromatiza-
tion, and dehydrogenation occurred to produce ethene, diethyl
ether, and acetaldehyde as primary products. The secondary prod-
ucts formed from ethene. The presence of cerium in the HZSM-5

b1469_Ch-13.indd 785 4/8/2013 12:41:51 PM


b1469 Catalysis by Ceria and Related Materials

786 L. Vivier and D. Duprez

zeolite (2.5, 5.0, 7.5, or 15 wt%) affected the acid properties of the
catalysts, and also improved the selectivity of C3 and C4 olefins. The
number of strong acid sites of the HZSM-5 zeolite decreased upon
cerium modification, which decreased the rate of coke formation
and enhanced the stability of the catalyst. The introduced cerium
promoted the formation of the C3 olefin from the metathesis of
ethene and cis-2-butene (converted rapidly from 1-butene), which
further increased the changes in the distribution of the C2–C4 ole-
fins compared with the HZSM-5 zeolite.

13.3 Synthesis of Carbonates


The development of environmental processes based on the utiliza-
tion of naturally abundant carbon resources such as carbon dioxide
has gained considerable attention in recent years. Organic carbon-
ate synthesis using carbon dioxide is one of the promising reactions
in this respect. Organic carbonate compounds have been used as
both a reactive intermediate and an inert solvent.
Dimethyl carbonate (DMC) can be synthesized from epoxide
compounds such as ethylene oxide or propylene oxide by a two-step
reaction (Fig. 13.1). In the first step the epoxide reacts with CO2,
producing a corresponding cyclic carbonate. In the second step, the
carbonate is transesterified with methanol to DMC and a corre-
sponding glycol.
CeO2 with several of the metal oxides does not appear to be a good
catalyst for this reaction. Basic metal-oxide catalysts have high activity
and selectivity in the reaction of epoxides and CO2 to the correspond-
ing cyclic carbonates and in transesterification with methanol. Of the
catalysts examined, MgO and CeO2 are the best catalysts and are active

O O

O CO2 O O CH3OH HO OH
O O
+
R step 1 st e p 2 CH3 CH3
R R

Figure 13.1 Two-step synthesis of dimethyl carbonate.

b1469_Ch-13.indd 786 4/8/2013 12:41:51 PM


b1469 Catalysis by Ceria and Related Materials

Transformation of Oxygenated Compounds 787

Table 13.3 Reaction of propylene epoxide (R = −CH3) with carbon


dioxide to cyclic carbonate (Step 1) and reaction of cyclic carbonate
and methanol to dimethyl carbonate (DMC) (Step 2).12
Stepa Catalyst MgO CeO2
1 Conversion of propylene oxide (%) 34.9 22.7
Selectivity of cyclic carbonate (%) 92.0 76.9
Yield of cyclic carbonate (%) 32.1 17.5
2 Conversion of cyclic carbonate (%) 28.0 32.8
Selectivity of dimethyl carbonate (%) 100 98.8
Yield of dimethyl carbonate (%) 28.0 32.4
a
Reaction conditions: Step 1: propylene oxide: 57 mmol; CO2: 8 MPa; catalyst: 1 g
(dimethylformamide, DMF: 4 ml); temperature: 150 °C; reaction time: 15 h.
Step 2: cyclic carbonate: 25 mmol; methanol: 200 mmol; CO2: 8 MPa; catalyst: 0.5 g
(DMF: 2 ml); temperature: 150 °C; reaction time: 4 h.

Table 13.4 Reaction of propylene carbonate (PC) with methanol to dimethyl


carbonate (DMC).14 Reaction conditions: PC: 10 mmol, MeOH: 100 mmol, catalyst:
115 mg; temperature: 140°C; reaction time: 6 h.
Catalyst PC conversion (%) DMC selectivity (%) DMC yield (%)
CeO2 33 30 10
0.08 wt% Au/CeO2 40 73 30
0.5 wt% Au/CeO2 63 55 35
1.5 wt% Au/CeO2 68 41 28

for both these reactions. However, MgO is preferred to CeO2 due to


lower cost and better selectivity for the first step (Table 13.3).12
Abimanyu et al. studied the synthesis of DMC by ethylene carbonate
transesterification with methanol (step 2) with various MgO–CeO2
mixed oxide catalysts at 150°C. They showed that a high ethylene car-
bonate conversion of 64% (DMC selectivity of 95%) could be obtained
by a catalyst with a cerium content around 25 mol%.13
Juárez et al. showed that ceria nanocrystallites are a moderately
active catalyst for the transesterification of propylene carbonate (PC)
by methanol to DMC at 140°C.14 The presence of gold nanoparticles
on the ceria at an appropriate loading significantly increased the
activity and selectivity towards transesterification (Table 13.4).

b1469_Ch-13.indd 787 4/8/2013 12:41:51 PM


b1469 Catalysis by Ceria and Related Materials

788 L. Vivier and D. Duprez

Tomishige et al. showed that CeO2–ZrO2 catalysts (with Ce/Ce+


Zr = 0, 0.2, 0.33, 0.5, and 1.0) are effective in the direct synthesis of
cyclic carbonate (ethylene carbonate (EC) or propylene carbonate
(PC)) from CO2 and diols (ethylene glycol and propylene glycol,
respectively) at various temperatures (110°C to 190°C).15,16 Catalytic
activity in the formation of PC and EC was very dependent on the
composition and calcination temperature of the CeO2–ZrO2 catalysts.
The active sites could be weak acid–base sites, which are present on the
plain surface of the catalysts calcined at high temperatures. The pro-
duction of EC and PC from the corresponding glycols and CO2 using
CeO2-based catalysts is limited by the reaction equilibrium.
Non-cyclic organic carbonates were also directly synthesized by a
reaction between methanol or ethanol and CO2 in the presence of
CeO2–ZrO2 catalysts17 or CeO2 catalysts.18–20 The CeO2–ZrO2 catalyst
(Ce/Ce+Zr = 0.2) seems to be very effective for the selective synthe-
sis of DMC from CH3OH and CO2. Although the selectivity of DMC
synthesis over the CeO2–ZrO2 catalyst was very high (100%) under
the reaction conditions, unfortunately methanol conversion was
very low because the equilibrium of the reaction was largely shifted
to the reactant.17
The synthesis of DMC from CH3OH and CO2 was investigated on
CeO2 prepared with various kinds of precursors under various calci-
nation temperatures. The BET surface area decreased with increas-
ing calcination temperature. One sample of CeO2 calcined at 600°C
with a BET surface area of 80 m2·g−1 exhibited the highest activity of
all the CeO2 catalysts. Overall, the formation rate of DMC was almost
proportional to the BET surface area of the catalysts. This suggests
that the active site of this reaction is on a stable crystal surface of
CeO2, such as (111).18
Methanol conversion was very low because the formation of
DMC was limited by the reaction equilibrium. However, when H2O
is removed from the reaction system, it is possible to enhance meth-
anol conversion drastically. H2O can be removed by reaction with
acetals such as 2,2-dimethoxypropane (DMP)21 using a CeO2–ZrO2
catalyst, or by reaction with acetonitrile22,23 or with benzonitrile24
catalyzed by CeO2.

b1469_Ch-13.indd 788 4/8/2013 12:41:51 PM


b1469 Catalysis by Ceria and Related Materials

Transformation of Oxygenated Compounds 789

The life of CeO2 catalysts used in the direct carboxylation of


methanol to dimethyl carbonate is quite short, as after the first cycle
the activity decreases and after a few cycles it goes to zero19,20 This
deactivation is mainly due to a surface modification produced by the
Ce(IV) to Ce(III) reduction during catalysis and to crystal conglom-
eration. The modification of ceria by loading alumina strongly
reduces the oxidation of methanol and the consequent reduction of
Ce(IV) to Ce(III) with an increase in both the life of the catalyst and
selectivity.
Recently, CeO2-based catalysts (Al2O3/CeO2 or Nb2O5/CeO2)
were used in the direct carboxylation of glycerol with CO2
(Table 13.5).25 This reaction is of great industrial interest: it converts
two waste materials into an added-value product, glycerol carbonate,
which is a bifunctional compound used as a solvent, a surfactant,
and in the synthesis of polyurethanes and polycarbonates.
Unfortunately, only 10% glycerol conversion was observed at 180°C.
No reaction was observed below 130°C. Previously, these authors26
had shown that for the glycerolysis of urea, in which urea is used as
an activated form of CO2, the conversion was slightly higher with
CeO2 catalysts than for a thermal process (without a catalyst).

13.4 Reaction of Glycerol


Glycerol is obtained from biomass via hydrolysis or methanolysis of
triglycerides. The growth in the production of biodiesel by the trans-
esterification of oil is responsible for the increased production of
glycerol and its lower cost. This could lead to many applications for
glycerol, which can be transformed into valuable chemicals.27
Figure 13.2 shows several different reaction pathways, which may
involve ceria-based catalysts.

Table 13.5 BET surface area determination and acid/base site ratio.25
Catalyst BET surface area (m2·g−1) Acid/base site ratio
Al2O3/CeO2 81 0.53
Nb2O5/CeO2 36 1.09

b1469_Ch-13.indd 789 4/8/2013 12:41:51 PM


b1469 Catalysis by Ceria and Related Materials

790 L. Vivier and D. Duprez

O OH OH OH
HO OH HO OH HO OH HO
O
O O O
dihydroxyacetone glyceric acid tartronic acid glycolic acid

oxidation

OH transesterification OH hydrogenolysis OH
HO OH
HO O R HO OH HO

O glycerol 1,2-propanediol ethylene


monoglycerides glycol
dehydration

O
OH O
hydroxyacetone acrolein

Figure 13.2 Transformation of glycerol on ceria-based catalysts.

13.4.1 Oxidation
For the production of valuable oxygenated derivatives, glycerol can be
oxidized forming a large number of products such as dihydroxyace-
tone (DHA), glyceric acid (GLYAC), tartronic acid (TARAC), glycolic
acid, or hydroxyethanoic acid (HEA) (Fig. 13.2) among many others.
These products can be used as intermediate compounds in the
synthesis of useful chemicals.
Ceria-supported catalysts are well known in oxidation reactions
and gold nanoparticles supported on ceria are excellent general
heterogeneous catalysts for the aerobic oxidation of alcohols (see
Vivier and Duprez28 and references therein). Corma and colleagues
showed that the efficiency of the catalysts can be improved by using
ceria supports with high specific surface areas and attributed this
behavior to a cooperative effect between gold and ceria.29–33
Demirel et al.34 studied the liquid-phase oxidation of polyalcohols
with both primary and secondary alcohol groups; they compared the
reactivity of n-propanol and glycerol at 60°C, pH = 12, pO2 = 1 atm on
ceria-supported gold catalysts. It was difficult to correlate oxidation
activity with the specific surface area: the most active catalyst was

b1469_Ch-13.indd 790 4/8/2013 12:41:51 PM


b1469 Catalysis by Ceria and Related Materials

Transformation of Oxygenated Compounds 791

1 wt% Au/CeO2 with a specific surface area equal to 60 m2·g−1. No


influence of the specific surface on selectivity could be detected (glyc-
eric acid: 45%; dihydroxyacetone: 25%; glycolic acid: 12%, and tar-
tronic acid: 8%). Unfortunately, ceria-supported gold catalysts are less
active than carbon supported gold catalysts.

13.4.2 Dehydration
Glycerol can be converted into hydroxyacetone and acrolein by
dehydration (Fig. 13.2). Acrolein is an important intermediate used
as a raw material in the production of acrylate polymers.
The vapor-phase dehydration of glycerol to hydroxyacetone was
also studied by Sato et al. on copper catalysts at 205°C.35 They found
that the acid–base property of the support material affected the
selectivity; basic MgO, CeO2, and ZnO supports showed low hydroxy-
acetone selectivity, while acidic supports, such as Al2O3, ZrO2, Fe2O3,
and SiO2, effectively promoted hydroxyacetone selectivity.
Kinage et al.36 showed that the 5 wt% Na/CeO2 catalyst is one of
the best catalysts for high glycerol conversion with high selectivity for
hydroxyacetone at 350°C (Table 13.6). Other by-products are propa-
noic acid, ethylene glycol, 1,2-propanediol and allyl alcohol, with a
selectivity below 10%. Glycerol conversion was carried out on a
fixed-bed continuous-flow reactor at 350°C. For all catalysts, glycerol
conversion decreased initially up to 8 h and then became stable.

Table 13.6 Base/acid site ratio of Na-doped metal-oxide catalysts and their cata-
lytic performance for the dehydration of glycerol.36
Base/acid Hydroxyacetone Conversion after Hydroxyacetone
Catalyst site ratio selectivity (%)a deactivation (%)b selectivity (%)c
Na/CeO2 2.1 69 25 75
Na/Al2O3 3.6 57 15 65
Na/Ga2O3 3.2 35 5 55
Na/ZrO2 2.9 50 5 35
a
Hydroxyacetone selectivity at 10% glycerol conversion.
b
Glycerol conversion after deactivation (Time on stream = 15 h).
c
Hydroxyacetone selectivity after deactivation (Time on stream = 15 h).

b1469_Ch-13.indd 791 4/8/2013 12:41:51 PM


b1469 Catalysis by Ceria and Related Materials

792 L. Vivier and D. Duprez

Table 13.7 Properties of oxide catalysts and their catalytic performance for the
dehydration of glycerol.37 Reaction conditions: temperature: 250°C; time on
stream: 8–10 h.
Surface area Total basic sites Hydroxyacetone
Oxide catalyst (m2·g−1) (mmol CO2·g−1) Conversion (%) selectivity (%)
Ce 260 0.36 35.6 1.2
Zr 192 0.05 61.3 39.8
Ti 402 0.29 70.2 31.0
CeZr 122 0.21 93.5 42.0

Table 13.8 Properties of ternary composites.38


Composition Surface area Total acidic sites Total basic sites
Ternary composite (%wt) (m2·g−1) (mmol NH3·g−1) (mmol CO2·g−1)
CeO2–ZrO2–Co3O4 71.6:7.4:16.2 113 0.25 0.75
CeO2–ZrO2–NiO 70.2:7.5:15.8 52 0.89 0.12
CeO2–NiO–Al2O3 73.0:8.7:17.9 87 1.02 0.07

The performance of various nanocasted oxides was studied by


Vasconcelos et al. in the gas-phase dehydration of glycerol to
produce hydroxyacetone (Table 13.7).37 The CeZr catalyst exhibited
the best catalytic performance, CeO2 being the worst. The binary
oxide CeO2–ZrO2, with moderate basicity, proved to have a coopera-
tive acid–base character in the Ce0.8Zr0.2O2 phase, resulting in a
higher production of hydroxyacetone.
Ternary composites have been prepared and have both basic and
acidic sites (Table 13.8).38 The CeO2–ZrO2–Co3O4 catalyst had the best
performance in terms of the highest acrolein and 1-hydroxyacetone
selectivity and conversion due to the Co3O4 phase stabilizing the nano-
particles of CeO2–ZrO2. Although CeO2–NiO–Al2O3 and CeO2–ZrO2–
NiO are both strong acids, which are active and selective in obtaining
acrolein, both were deactivated after 6 h under the reaction condi-
tions of 250 °C due to physical degradation of the catalysts.38
The dehydration of glycerol to acrolein has been carried out in
the gas phase using various solid catalysts with a wide range of acid–
base properties.39 The solid base catalysts were characterized by their

b1469_Ch-13.indd 792 4/8/2013 12:41:51 PM


b1469 Catalysis by Ceria and Related Materials

Transformation of Oxygenated Compounds 793

low selectivity for the production of acrolein. The main products


identified were 1,2-propanediol over the CeO2 catalyst. It seems that
hydrogenation of the carbonyl group in hydroxyacetone, which pro-
duces 1,2-propanediol, was only possible when the solid bases were
used as the catalysts. Strongly acidic catalysts showed higher acrolein
selectivity than catalysts with weaker acidity.
Similar results were obtained in the gas-phase dehydration of
glycerol to acrolein over 10 wt% HSiW catalysts supported on differ-
ent supports, γ-Al2O3, SiO2–Al2O3, TiO2, ZrO2, SiO2, AC, CeO2, and
MgO.40 For comparison, the same reaction was conducted over each
support without HSiW. Ceria showed the highest hydroxyacetone
selectivity at 315°C of the various metal oxides. Glycerol conversion
to acrolein generally increased with an increasing number of acid
sites. HSiW/SiO2–Al2O3, HSiW/γ–Al2O3, and HSiW/ZrO2 were the
best catalysts (Table 13.9). When CeO2 was used as support, acrolein
selectivity was very low. It is clear that the basic sites are not effective
for the dehydration of glycerol to acrolein.
Similar results were obtained with a series of MeO–Al2O3–PO4
catalysts prepared to evaluate the effect of different transition-metal
oxides loaded on phosphated alumina as catalysts for the gas-phase
dehydration of glycerol at 280°C. For the CeOx–Al2O3–PO4 catalyst,
the hydroxyacetone selectivity is higher than in acrolein.41

13.4.3 Hydrogenolysis
Of the various promoters, cerium showed excellent promotion
effects in hydrogenation reactions (see Vivier and Duprez28 and

Table 13.9 Properties of catalysts and their catalytic performance for the dehydra-
tion of glycerol.40 Reaction conditions: temperature: 315°C; time on stream: 2 h.
Surface area Glycerol Acrolein Hydroxyacetone
Oxide catalyst (m2·g−1) conversion (%) selectivity (%) selectivity (%)
HSiW/SiO2–Al2O3 465 100.0 58.0 9.6
HSiW/ZrO2 66 83.6 58.1 11.9
HSiW/γ–Al2O3 218 87.0 45.5 13.7
HSiW/CeO2 248 62.4 5.1 23.4

b1469_Ch-13.indd 793 4/8/2013 12:41:52 PM


b1469 Catalysis by Ceria and Related Materials

794 L. Vivier and D. Duprez

references therein). Yu et al. described a procedure for preparing


a Ni/AC catalyst with activated carbon (AC) as support for
aqueous glycerol hydrogenolysis.42 The addition of cerium to the
Ni/AC catalyst (BET surface area: 957 m2·g−1; total acidic sites:
0.24 mmol NH3·g−1) showed a remarkable promotion effect on
catalytic performance. Glycerol was converted at 90.4% at 200°C
under 5 MPa of H2 with 65.7% selectivity of 1,2-propanediol and
10.7% selectivity of ethylene glycol. The addition of cerium to
the Ni/AC catalyst greatly changed the reductive behavior.
Glycerol hydrogenolysis was possible according to the proposed
dehydration/hydrogenation mechanism. Glycerol was dehydrated
to acetol and then hydrogenated to 1,2-propanediol as shown by
Chai et al.39

13.4.4 Transesterification
Monoglycerides, the glycerol monoesters of fatty acids, are biode-
gradable and non-toxic molecules with a hydrophilic head and a
hydrophobic tail, having surfactant and emulsifying properties.
They can be used in foods, detergents, plasticizers, and cosmetic
and pharmaceutical formulations.43
The preparation of monoglycerides from fatty acids or fatty
methyl esters and glycerol can be carried out in the presence of
acidic or basic catalysts. The use of solid basic catalysts could limit
secondary reactions leading to product degradation. A comparison
of various basic oxide solids has shown that the more significant
the intrinsic basicity, the more active the catalyst is for the transes-
terification of methyl stearate with glycerol (Table 13.10).44,45
Barrault and colleagues showed that basic and solid catalysts such
as MgO and CeO2 can easily replace homogeneous catalysts while
having the same selectivity. Similar results were obtained for the
transesterification of methyl oleate with glycerol at 220°C with
MgO, Y2O3, and CeO2 catalysts. These catalysts converted more
than 90% of the methyl oleate with about 70% of monoglyceride
selectivity.46

b1469_Ch-13.indd 794 4/8/2013 12:41:52 PM


b1469 Catalysis by Ceria and Related Materials

Transformation of Oxygenated Compounds 795

Table 13.10 Catalyst properties and their catalytic performance for the transes-
terification of methyl stearate with glycerol.44,45 Reaction conditions: temperature:
220°C; glycerol/methyl stearate: 1; time on stream: 6 h; weight of catalyst: 0.5 g; N2
atmospheric pressure.
Oxide BET surface Total acidic sites Total basic sites Conversion Monoglyceride
catalyst area (m2·g−1) (µmol NH3·g−1) (µmol CO2·g−1) (%) selectivity (%)

ZnO 33 455 21 18 80
MgO (I) 13 110 202 83 42
MgO (II) 151 290 345 80 38
La2O3 5 565 80.6 97 28
CeO2 (I) 19 546 16.3 4.1 100
CeO2 (II) 135 805 345 82 42

13.5 Selective Hydrogenation of C=O Bonds


of α,β-Unsaturated Aldehydes
The selective hydrogenation of α,β-unsaturated aldehydes to unsatu-
rated alcohol is an important reaction in the production of many
pharmaceutical, agrochemical, and fragrance compounds. The
hydrogenation of the C=C bond is thermodynamically more favora-
ble than C=O hydrogenation, and low yields of the desired product
are obtained with conventional hydrogenation catalysts.
Cerium-based platinum catalysts have been extensively studied
for the hydrogenation of crotonaldehyde (CH3−CH=CH−CHO)47–58
or citral ((CH3)2C=CH−(CH2)2−C(CH3)=CH−CHO).59,60 The activa-
tion of the carbonyl bond is induced by the presence of oxygen
vacancy sites located at the interface between ceria and the platinum
particles.
The selective hydrogenation of α,β-unsaturated aldehydes is
used as a probe reaction in studying the strong metal/support inter-
action (SMSI). Ceria is able to form oxygen vacancies and interme-
tallic compounds after reduction treatment at relatively high
temperatures.
Touroude and collaborators studied the selective hydrogenation
of crotonaldehyde on 5 wt% Pt/CeO2 in the gas phase at atmospheric
pressure.47–49 The properties of the catalysts and their catalytic

b1469_Ch-13.indd 795 4/8/2013 12:41:52 PM


b1469 Catalysis by Ceria and Related Materials

796 L. Vivier and D. Duprez

Table 13.11 Catalyst properties and catalytic performance in the hydrogenation


of crotonaldehyde.47,48
Catalyst A Catalyst B Catalyst C
Precursor Pt(NH3)4(NO3)2 H2PtCl6 Pt(NH3)4Cl2
Reduction (°C)a 200 700 200 700 200 700
2 −1
Surface area (m ·g ) 54.6 36.9 54.6 51.7 54.6 52.7
Cl content (wt%) — — 1.5 1.0 1.1 1.0
−1
Catalytic activity (µmol(s·gPt) ) 24.8 2.1 55.4 12 30.3 11.7
Butanal selectivity (%) 85.6 9.5 88.6 65.3 87.7 70.6
Crotyl alcohol selectivity (%) 7.5 83.0 3.7 15.2 3.1 19.4
a
After calcination at 400°C for 4 h.

performance are shown in Table 13.11. On chlorine-free Pt/CeO2,


crotyl alcohol selectivity increased by up to more than 80% (45% con-
version) when the reduction temperature of the catalysts reached
700°C. The presence of chlorine, during catalyst synthesis, preserved
the catalytic properties of platinum metal for the hydrogenation of the
C=C bond; chlorine atoms around the platinum particles inhibit the
diffusion of cerium atoms inside the metal particles and prevent
the formation of the CePt5 alloy. By controlling the nanostructure,
size, and morphology of the supported platinum particles, the authors
showed that it was possible to orientate the selectivity in the hydro-
genation of crotonaldehyde.49
The presence of zinc facilitated the reduction of surface ceria,
thereby increasing the potential number of Ce3+–Pt metal interface
sites, particularly in cases where small Pt particles were located in
ceria-rich zones of the support.52,53,55,60 In the case of the vapor-phase
hydrogenation of crotonaldehyde, the overall catalytic activity increased
significantly after reduction at 500°C for the zinc-containing cata-
lyst, and furthermore, the selectivity toward the hydrogenation of
the carbonyl bond was improved (Table 13.12).
Pt on mesostructured CeO2 nanoparticles embedded within
ultrathin layers of a highly structured SiO2 binder had the highest
activity with 80% selectivity for the chemoselective hydrogenation
of crotonaldehyde.54 By increasing the reduction temperature, the

b1469_Ch-13.indd 796 4/8/2013 12:41:52 PM


b1469 Catalysis by Ceria and Related Materials

Transformation of Oxygenated Compounds 797

Table 13.12 Catalytic activity and selectivity over Pt/CeO2–SiO2 and


PtZn/CeO2–SiO2 for the hydrogenation of crotonaldehyde.53
Catalyst Pt/CeO2–SiO2 PtZn/CeO2–SiO2
Reduction (°C)a 200 500 200 500
Conversion (%) 10 8 2 27
Butanal selectivity (%) 92 56 74 63
Crotyl alcohol selectivity (%) 2 40 22 15
a
After calcination at 400°C for 4 h. Reaction conditions: temperature: 80°C;
time on stream: 4 min; H2/aldehyde ratio: 26; weight of catalyst: 100 mg; at
atmospheric pressure.

number of Pt–CeO2−x interfacial sites, which are responsible for


activating the carbonyl bond, increased.
The hydrogenation of citral has been carried out in the liquid
phase at 50°C in ethanol59 or at 70°C in isopropanol.60 The forma-
tion of geraniol (E isomer) has been observed as the sole product
on 5 wt% Pt/CeO2 (35 m2.g−1) and has been attributed to the influ-
ence of the SMSI state on the selective hydrogenation of the C=O
bond.59 On a 2 wt% Pt/CeO2/C catalyst (CeO2 loading of 20 wt%),
the main products of the hydrogenation of citral were citronellal
(hydrogenation of the conjugated C=C bond), the unsaturated alco-
hols geraniol and nerol (hydrogenation of the C=O bond), and the
alcohol citronellol (by hydrogenation of the C=O bond of citronel-
lal).60 On the one hand, the creation of new Pt–CeOx sites at the
metal/support interface act as Lewis acid sites able to activate the
C=O bond of the citral molecule; on the other hand, the electronic
interaction between the reduced ceria particles and the active metal
leads to an increase in electron density on the platinum particles,
with subsequent weakening of the adsorption of citral via the C=C
bond. Moreover, when tin was added (Sn/Pt atomic ratio varying
from 0 to 0.75), ceria reducibility increased. The presence of Snn+
species, also able to act as Lewis acid sites, on the surface of plati-
num particles or in their close vicinity could account for the increase
in selectivity for unsaturated alcohols with reduction temperature
(Table 13.13). The increase of conversion after reduction at high

b1469_Ch-13.indd 797 4/8/2013 12:41:52 PM


b1469 Catalysis by Ceria and Related Materials

798 L. Vivier and D. Duprez

Table 13.13 Catalyst properties and catalytic performance for the hydrogenation
of citral.60 Reaction conditions: temperature: 70°C; H2 pressure: 7 MPa; time on
stream: 60 min; weight of catalyst: 500 mg; 3 vol% citral/isopropanol.
Surface area Selectivity to
Catalyst (m2·g−1) Conversion (%) unsaturated alcohol (%)
Reduction (°C) 200 500 200 500
Pt/CeO2/C 428 95 75 8 16
PtSn(0.25)/CeO2/C 366 45 60 14 32
PtSn(0.5)/CeO2/C 363 40 32 18 35
PtSn(0.75)/CeO2/C 348 22 30 22 33

temperatures in these catalysts could also be explained by the crea-


tion of new Pt–Snn+ sites active for hydrogenation of the C=O bond
in the citral molecule.
The use of cerium on Ru-based catalysts supported on alumina
and activated carbon increased selectivity to unsaturated alcohols in
the hydrogenation of crotonaldehyde (in the gas phase) and citral
(in the liquid phase with isopropanol as solvent).61 However, the
observed activities were very low.
More recently, Campo et al. studied the influence of the specific
surface area of the support on the selective hydrogenation of cro-
tonaldehyde on Au/CeO2.62–65 Hydrogenation was carried out
either at 120°C and at H2 atmospheric pressure62–64 or at 80°C in
the liquid phase (isopropanol was used as a solvent).64,65 They
showed that a high-surface-area catalyst (Au/HAS–CeO2, with
240 m2.g−1) was active and highly selective to crotyl alcohol. Other
Au/CeO2 catalysts with lower specific surface areas showed rather
low selectivity (Table 13.14). High selectivity is an intrinsic charac-
teristic of gold particles (particle size lower than 4 nm on Au/HAS–
CeO2 and higher than 9 nm on other Au/CeO2 catalysts), though
ceria plays an important role as a result of its redox and acid–base
properties.
Recently, He et al. showed that gold supported on ceria can cata-
lyze the reduction of crotonaldehyde to the corresponding unsatu-
rated alcohol (crotyl alcohol) in high yields in the presence of CO

b1469_Ch-13.indd 798 4/8/2013 12:41:52 PM


b1469 Catalysis by Ceria and Related Materials

Transformation of Oxygenated Compounds 799

Table 13.14 Catalytic properties of Au/CeO2 catalysts


for the hydrogenation of crotonaldehyde under liquid
phase at 80°C at 240 min of reaction time.65
Catalyst Au/CeO2
Surface area (m2·g−1) 80 150 240
Conversion (%) 55 58 70
Crotyl alcohol selectivity (%) 10 20 29

and H2O.66 Compared with conventional catalytic hydrogenation


processes, the Au-catalyzed CO/H2O-mediated reduction system has
significant advantages such as requiring only mild conditions and
not directly needing molecular hydrogen. Au/CeO2 catalysts showed
the highest activity and selectivity for unsaturated alcohol. When
molecular hydrogen was applied instead of CO/H2O, a significant
decrease of the reduction rate (96% to 60% for the conversion of
crotonaldehyde) and crotyl alcohol selectivity (85% to 38%) was
observed.
The first selective hydrogenation of cinnamaldehyde over Pd
(2 wt%) supported on CeO2, ZrO2, and CeO2–ZrO2 catalysts was
described by Bhogeswararao and Srinivas.67 On these catalysts, the
conjugated C=C hydrogenation of cinnamaldehyde was preferred to
C=O hydrogenation to yield selectively the saturated aldehyde,
hydrocinnamaldehyde (C6H5−(CH2)2−CHO).
Selectivity to the unsaturated alcohol is governed by different
factors: the nature of the active metal, metal particle size, support
effects, and the presence of promoters or bimetallic phases.

13.6 C–C Coupling Reactions


13.6.1 Ketonization
Carboxylic acids and esters are common intermediate products
formed in biomass conversion processes and can be converted to
valuable products by ketonization. This reaction is a green and
atom-economical process because non-polluting by-products are
formed and no solvent or other elaborate reagents are necessary.

b1469_Ch-13.indd 799 4/8/2013 12:41:52 PM


b1469 Catalysis by Ceria and Related Materials

800 L. Vivier and D. Duprez

Symmetric and nonsymmetric ketones have been obtained by the


ketonization of carboxylic acids (Eq. 13.1) or ethyl esters (Eq. 13.2)
in the gas phase over ceria-based catalysts under flowing conditions:

R′COOH + RCOOH → R′COR + CO2 + H2O (13.1)


2 RCOOC2H5 → RCOR + CO2 + H2O + 2 C2H4 (13.2)

Usually, ketonization is carried out in the presence of ceria-based


catalysts (10–20% CeO2 supported on SiO2, TiO2, or Al2O3) between
300 and 450°C (see Vivier and Duprez28 and references therein).
The condensation of acetic acid to acetone has been carried out
over pure CeO2 from 300°C.68,69 Acetic acid conversion was almost
complete. The bulk structure of CeO2 (13.2 m2·g−1) was stable dur-
ing ketonization: acetic acid conversion was 51.3% and the selectiv-
ity to acetone was over 99.9%.69 The catalytic sites were suggested to
be Lewis acid–base pair sites, with the Lewis acid sites (Ce4+) being
reducible.68 Stubenrauch et al. had shown previously using TPD that
acetone is produced during the decomposition of acetic acid on the
CeO2 (111) surface only.70
A Ce0.5Zr0.5O2 catalyst showed desirable catalytic properties for
the ketonization of carbohydrate-derived carboxylic acids in the
presence of other monofunctional oxygenated species, such as alco-
hols or ketones.71 Gaertner et al. developed a kinetic model that
describes the ketonization of mixtures containing carboxylic acids
and esters over Ce0.5Zr0.5O2 catalysts from 175 to 350°C.72–74 Under
these conditions two different reactions take place, esterification
and ketonization, both consuming hexanoic acid, used as a model
molecule. Ester ketonization is slower than the ketonization of the
corresponding acid.
Dumesic and collaborators developed a process for the catalytic
upgrading of levulinic acid to 5-nonanone, with the intermediate
formation of γ-valerolactone (Fig. 13.3).75,76 Levulinic acid can be
obtained inexpensively and in high yields via acid hydrolysis of waste
cellulosic materials. The process starts with the deconstruction of
solid cellulose in an aqueous solution of sulfuric acid yielding an
equimolar mixture of levulinic acid and formic acid. The formic

b1469_Ch-13.indd 800 4/8/2013 12:41:52 PM


b1469 Catalysis by Ceria and Related Materials

Transformation of Oxygenated Compounds 801

O
O O O
OH

O
levulinic acid γ-valerolactone 5-nonanone

Figure 13.3 Transformation of levulinic acid to 5-nonanone.

acid in this mixture can then be used to reduce levulinic acid to


γ-valerolactone in the sulfuric acid solution over a Ru/C catalyst.
This γ-valerolactone product, with residual amounts of sulfur, can be
upgraded to 5-nonanone with high yields (90%) in a single reactor
by using a dual catalyst bed of Pd/Nb2O5 plus Ce0.5Zr0.5O2.
Long-carbon-chain ketones (C17H35COC17H35, C15H31COC15H31,
CH3COC17H35, CH3COC15H31) have also been obtained from methyl
esters of fatty acids (essentially C17H35COOCH3 and C15H31COOCH3)
in methanol at atmospheric pressure at 385 °C (the optimal tempera-
ture), over catalysts containing Sn–Ce–Rh oxides in a molar ratio
90:9:1 (total yield: 63%, conversion: 96%).77 A similar catalyst was
used to transform methyl laurate (C11H23COOCH3) to 12-tricosanone
(C11H23COC11H23).78
Novel catalysts were synthesized by generating uniform particles
of cerium and manganese oxides (MnOx/CeO2) in situ within hex-
agonal mesoporous silica (CM-HMS) and MCM-41 (CM-MCM-41)
supports (Table 13.15).79
The conversions obtained over CM-HMS and CM-MCM-41 are
similar to those measured over unsupported MnOx/CeO2. Thus, the
catalytic activity of these catalysts for the ketonization of carboxylic

Table 13.15 Catalyst properties and catalytic performance for the ketonization of
propanoic acid (PA) or butanoic acid (BA).79 Reaction conditions: temperature:
410 °C; N2 flow: 15 cm3.min−1; acid feed rate: 0.1 cm3.min−1; weight of catalyst: 1.5 g.
Surface area Total basic sites Conversion Conversion
Catalyst (m2·g−1) (µmol·g−1) of PA (%) of BA (%)
MnOx/CeO2 106 32 80 28
CM-HMS 250 26 78 36
CM-MCM-41 529 28 73 31

b1469_Ch-13.indd 801 4/8/2013 12:41:52 PM


b1469 Catalysis by Ceria and Related Materials

802 L. Vivier and D. Duprez

acids was confirmed; better utilization of Ce and Mn was observed


than in unsupported MnOx/CeO2 because CM-HMS and CM-MCM-41
contain only 46 wt% MnOx/CeO2.
Ketones can also be produced by the dimerization of alcohols
over ceria-based catalysts (see Vivier and Duprez28 and references
therein) by a reaction mechanism in which alcohol is oxidized to
aldehyde and then carboxylic acid following by the coupling of two
equivalent acids to give a symmetrical ketone and CO2.

13.6.2 Aldol condensation


The catalytic conversion of ethanol and glycerol produces short alde-
hydes or ketones, which can be converted to more suitable molecular
weight products. Aldol condensation reactions of these molecules
lead also to the formation of C–C bonds. These reactions are gener-
ally catalyzed by bases. These types of condensation over solid bases
lead mainly to aldol, ketol, or α,β-unsaturated carbonyl compounds.
The aldol condensation of acetone has been studied over solid
base catalysts such as Ca(OH)2, La(OH)3, ZrO2, and CeO2 in the
vapor phase between 200 and 400 °C. The condensation of acetone
(Fig. 13.4, compound 1) gives diacetone alcohol (compound 2),
which is dehydrated to mesityl oxide (compound 3). Various sec-
ondary products are formed by numerous secondary reactions, such
as further aldolization and Michael condensation. At 200–400°C,
compound 1 over CeO2 had 74–97% conversion. CeO2 promoted
the formation of 1,3,5-trimethylbenzene (compound 7) with a selec-
tivity of 47% at 400°C, but also produced large amounts of higher
condensation products (38.9% at 300°C). The formation of com-
pound 7 is favored with decreasing basic strength. Indeed, the basic-
ity and basic strength, respectively, of the catalysts decreased in the
order Ca(OH)2 > La(OH)3 > CeO2 > ZrO2.80
The reaction of acetaldehyde has also been studied on CeO2-
based catalysts. CeO2 was chosen as a support because its reducibility
and basicity favor aldolization reactions.81,82 Three C–C bond forma-
tion reactions from acetaldehyde were observed: aldolization to cro-
tonaldehyde and crotyl alcohol (more prominent on CeO2 alone),

b1469_Ch-13.indd 802 4/8/2013 12:41:52 PM


b1469 Catalysis by Ceria and Related Materials

Transformation of Oxygenated Compounds 803

Figure 13.4 Condensation of acetone and secondary reactions.

ketonization to acetone, and reductive coupling to form butenes


and butadiene.81
Methyl isobutyl ketone (MIBK) or 4-methyl-2-pentanone was
synthesized from 2-propanol in one pot on bifunctional metal/acid–
base catalysts. The synthesis of MIBK from 2-propanol involves the
dehydrogenation of 2-propanol to acetone, which is converted to
mesityl oxide (compound 3, Fig. 13.4) via an aldol condensation
reaction and consecutive dehydration of the aldol intermediate,
diacetone alcohol (compound 2, Fig. 13.4). Mesityl oxide is hydro-
genated on the metallic site to MIBK by H2 generated during 2-pro-
panol dehydrogenation. One of the most selective catalysts for the
formation of MIBK is CuCe4Ox (with CuMg10Al7Ox), which has a
higher density of base sites and a lower density of acid sites than a
CuAl16Ox catalyst (Table 13.16).83,84
In order to transform the oxygenated compounds derived from
biomass into valuable compounds, Dumesic and collaborators85,86
studied the aldol condensation/hydrogenation reaction of 2-hex-
anone (as A model molecule) over a Pd/CeZrOx catalysts at 350°C
under 5 bar (Table 13.17).
The primary product of aldol condensation/hydrogenation is
C12 ketone, with the formation of C9 and C18 ketones as second-
ary products. The selectivity towards C12 products (essentially

b1469_Ch-13.indd 803 4/8/2013 12:41:52 PM


b1469 Catalysis by Ceria and Related Materials

804 L. Vivier and D. Duprez

Table 13.16 Catalyst properties and selectivity to MIBK from 2-propanol.84


Reaction conditions: temperature: 200 °C; H2 atmospheric pressure; 2-propanol
partial pressure: 7.8 kPa; 2-propanol conversion: 40%.
Surface area Total acidic sites Total basic sites Selectivity to
Catalyst (m2·g−1) (µmol·m−2) (µmol·m−2) MIBK (%)
CuAl16Ox 211 1.2 0.1 1.9
CuMg10Al7Ox 211 0.5 1.6 11.2
CuCe4Ox 74 1.0 2.3 10.2
CuMg10Ce2Ox 102 0.3 2.7 7.6
CuMg10Ox 150 0.1 5.0 6.1

Table 13.17 Catalyst properties and catalytic performance with different metal
loading85 and with different Ce/Zr ratios.86 Reaction conditions: temperature:
350 °C; H2 pressure: 5 bar; weight of catalyst: 2 g; weight hourly space velocity
(WHSV): 1.92 h−1.
Surface area Basic sites Acidic sites Conversion Selectivity
Catalyst (m2·g−1) (µmol CO2·g−1) (µmol NH3·g−1) (%) to C12 (%)
CeZrOx 130 370 56 21 66
0.05% Pd/CeZrOx 135 390 50 68 85
0.25% Pd/CeZrOx 133 380 54 58 83
1.45% Pd/CeZrOx 139 395 77 85 57
0.25% Pd/CeOx 88 266 0 40 77
0.25% Pd/Ce5Zr2Ox 131 414 27 58 79
0.25% Pd/Ce1Zr1Ox 133 380 54 58 83
0.25% Pd/Ce2Zr5Ox 141 350 86 84 77
0.25% Pd/ZrOx 149 296 212 90 57

7-methyl-5-undecanone with the presence of 5-methylundecane)


decreases with increasing metal loadings.85 The 2-hexanone conver-
sion increases with increasing ZrO2 content, pure ZrO2 having the
highest conversion of 90%. Selectivity towards the C12 products
remains approximately constant for all catalysts except for pure
ZrO2.86 The CeZrOx support was selected because it possesses high
lattice oxygen mobility and because of its ability to interact strongly
with supported metals. Previously, these authors found that 0.25 wt%

b1469_Ch-13.indd 804 4/8/2013 12:41:52 PM


b1469 Catalysis by Ceria and Related Materials

Transformation of Oxygenated Compounds 805

Pd/CeZrOx is stable at 350 °C for the aldol condensation of the com-


plex mixture produced by the transformation of glucose in aqueous
solution on 10 wt% Pt–Re/C at 210°C under 18 bar.71
A CexZr1-xO2 catalyst has been used for propanal condensation at
400°C under He or H2.87 Propanal is converted by two pathways, aldol
condensation and ketonization, to 3-pentanone, 2-methyl-2-pentenal,
2-methylpentanal, 3-heptanone, and 4-methyl-3-heptanone.

13.7 Oxidation of Sugar Derivatives


Gold nanoparticles supported on nanocrystalline cerium oxide, are
extremely effective in the oxidation of aldehydes to acids and also in
the oxidation of alcohols (see Vivier and Duprez28 and references
therein).88,89
A key intermediate in the synthesis of a number of pharmaceu-
ticals, antifungals, and polymer precursors is 5-hydroxymethyl-2-fur-
fural (HMF), which is produced by the dehydration of fructose or
glucose. Of these applications, the synthesis of polymer precursors
is of interest. For example, 2,5-furandicarboxylic acid (FDCA) can
replace terephthalic acid commonly used in the manufacture of
poly(ethylene terephthalate) (PET).
HMF has been selectively converted into FDCA (99 mol% yield)
in water, under mild conditions (65–130°C, 10 bar air) using gold
nanoparticles on nanoparticulated ceria.90 As an alternative to
FDCA, 2,5-dimethylfuroate (FDMC) has also been synthesized using
the same catalyst in the absence of a base in methanol.91 The oxida-
tion of HMF into FDCA or FDMC comprises two steps: aldehyde
oxidation and alcohol oxidation. Kinetic studies show that the rate-
limiting step of the reaction is alcohol oxidation to aldehyde. Once
the aldehyde is formed, the corresponding hemiacetal is obtained,
which is rapidly oxidized into the acid or the ester (Fig. 13.5).
The selective oxidation of arabinose to arabinonic acid by
molecular oxygen has been carried out at 60°C in water on mono-
metallic and bimetallic Pd–Au catalysts supported on nanosized
ceria.92 Arabinose is extracted from a hemicellulose called arabi-
nogalactan found in larch species, with galactose and D-glucuronic

b1469_Ch-13.indd 805 4/8/2013 12:41:52 PM


b1469 Catalysis by Ceria and Related Materials

806 L. Vivier and D. Duprez

HO HO HO OH
O O O
H2 O O OH O O O O
FDCA

O
O OH
HMF

CH3OH O O O O
O O O
O OH O O O O
FDMC

Figure 13.5 Reaction pathway for HMF transformation to FDCA or FDMC.

acid. Bimetallic Pd–Au catalysts have higher activity and selectivity in


comparison with monometallic gold and palladium samples. PdO
and Au species form core–shell structures while interactions of Au
and Pd with reducible ceria coexist with the mutual interaction
between these metals.

13.8 Conclusions
In recent years, increasing effort has been spent in finding ways to
use biomass as a feedstock for the production of organic chemicals.
The conversion of biomass can be achieved through biochemical or
thermochemical processes to obtain a wide range of oxygenated
compounds such as aldehydes, ketones, alcohols, carboxylic acids,
and esters. These products can be a source of valuable chemicals.
Pure ceria or ceria-based catalysts can be used in several organic
reactions, such as the dehydration and dehydrogenation of alcohols,
carbonatation, transesterification, aldolization, ketonization of car-
boxylic acids and ester, and in redox reactions.
The redox ability and the acid–base properties of CeO2, either
alone or in the presence of transition metals, are important parame-
ters that activate complex organic molecules and selectively orient
their transformation. The acid–base or redox properties of ceria can
also be modified by involving other oxides (ZrO2, La2O3, MnOx, ZnO,
MoO4, VOx, …) thus increasing the scope of the reactions. Ceria has
been used as a support of noble metal catalysts (Pd, Pt, Rh, Au) in the

b1469_Ch-13.indd 806 4/8/2013 12:41:52 PM


b1469 Catalysis by Ceria and Related Materials

Transformation of Oxygenated Compounds 807

dehydrogenation of alcohol or the selective hydrogenation of α,β-


unsaturated aldehydes (to unsaturated or saturated alcohols).

References
1. Bernal, S., Kaspar, J., Trovarelli A., Catal. Today, 20 (1999) 173.
2. Trovarelli, A., Catalysis by Ceria and Related Compounds, Imperial
College Press, London, (2002).
3. Binet, C., Daturi, M., Lavalley, J.-C., Catal. Today, 50 (1999) 207–225.
4. Yee, A., Morrison, S.J., Idriss, H., J. Catal., 186 (1999) 279–295.
5. Yee, A., Morrison, S.J., Idriss, H., J. Catal., 191 (2000) 30–45.
6. Yee, A., Morrison, S.J., Idriss, H., Catal. Today, 63 (2000) 327–335.
7. Sheng, P.-Y., Bowmaker, G.A., Idriss, H., Appl. Catal. A: Gen., 261 (2004)
171–181.
8. Raskó, J., Kiss, J., Appl. Catal. A: Gen., 287 (2005) 252–260.
9. Nair, H., Gatt, J.E., Miller, J.T., Baertsch, C.D., J. Catal., 279 (2011)
144–154.
10. Zonetti, P.C., Celnik, J., Letichevsky, S., Gaspar, A.B., Appel, L.G.,
J. Mol. Catal. A: Chem., 334 (2011) 29–34.
11. Bi, J., Liu, M., Song, C., Wang, X., Guo, X., Appl. Catal. B: Environm.,
107 (2011) 68–76.
12. Bhanage, B.M., Fujita, S.-I., Ikushima, Y., Arai, M., Appl. Catal. A: Gen.,
219 (2001) 259–266.
13. Abimanyu, H., Soo Kim, C., Sung Ahn, B., Sang Yoo, K., Catal. Lett.,
118 (2007) 30–35.
14. Juárez, R., Corma, A., García, H., Green Chem., 11 (2009) 949–952.
15. Tomishige, K., Yasuda, H., Yoshida, Y., Nurunnabi, M., Li, B.,
Kunimori, K., Catal. Lett., 95 (2004) 45–49.
16. Tomishige, K., Yasuda, H., Yoshida, Y., Nurunnabi, M., Li, B.,
Kunimori, K., Green Chem., 6 (2004) 206–214.
17. Tomishige, K., Furusawa, Y., Ikeda, Y., Asadullah, M., Fujimoto, K.,
Catal. Lett., 76 (2001) 71–74.
18. Yoshida, Y., Arai, Y., Kado, S., Kunimori, K., Tomishige, K., Catal.
Today, 115 (2006) 95–101.
19. Aresta, M., Dibenedetto, A., Pastore, C., Cuocci, C., Aresta, B.,
Cometa, S., De Giglio, E., Catal. Today, 137 (2008) 125–131.

b1469_Ch-13.indd 807 4/8/2013 12:41:52 PM


b1469 Catalysis by Ceria and Related Materials

808 L. Vivier and D. Duprez

20. Aresta, M., Dibenedetto, A., Pastore, C., Angelini, A., Aresta, B.,
Pápai, I., J. Catal., 269 (2010) 44–52.
21. Tomishige, K., Kunimori, K., Appl. Catal. A: Gen., 237 (2002) 103–109.
22. Honda, M., Suzuki, A., Noorjahan, B., Fujimoto, K.-I., Suzuki, K.,
Tomishige, K., Chem. Commun., (2009) 4596–4598.
23. Honda, M., Kuno, S., Begum, N., Fujimoto, K.-I., Suzuki, K.,
Nakagawa, Y., Tomishige, K., Appl. Catal. A: Gen., 384 (2010) 165–170.
24. Honda, M., Kuno, S., Sonehara, S., Fujimoto, K.-I., Suzuki, K.,
Nakagawa, Y., Tomishige, K., ChemCatChem, 3 (2011) 365–370.
25. Dibenedetto, A., Angelini, A., Aresta, M., Ethiraj, J., Fragale, C.,
Nocito, F., Tetrahedron, 67 (2011) 1308–1313.
26. Aresta, M., Dibenedetto, A., Nocito, F., Ferragin, C., J. Catal., 268
(2009) 106–114.
27. Zhou, C., Beltramini, J.N., Fan, Y.-X., Lu, G.Q., Chem. Soc. Rev., 37
(2008) 527–549.
28. Vivier, L., Duprez, D., ChemSusChem, 3 (2010) 654–678.
29. Abad, A., Concepción, P., Corma, A., García, H., Angew. Chem. Int. Ed.,
44 (2005) 4066–4069.
30. Abad, A., Almela, C., Corma, A., García, H., Tetrahedron, 62 (2006)
6666–6672.
31. Abad, A., Almela, C., Corma, A., García, H., Chem. Commun., (2006)
3178–3180.
32. Abad, A., Corma, A., García, H., Pure Appl. Chem., 79 (2007)
1847–1854.
33. Abad, A., Corma, A., García, H., Chem. Eur. J., 14 (2008) 212–222.
34. Demirel, S., Kern, P., Lucas, M., Claus, P., Catal. Today, 122 (2007)
292–300.
35. Sato, S., Akiyama, M., Takahashi, R., Hara, T., Inui, K., Yokota, M.,
Appl. Catal. A: Gen., 347 (2008) 186–191.
36. Kinage, A.K., Upare, P.P., Kasinathan, P., Kyu Hwang, Y., Chang, J.-S.,
Catal. Commun., 11 (2010) 620–623.
37. Vasconcelos, S.J.S., Lima, C.L., Filho, J.M., Oliveira, A.C., Barros, E.B.,
de Sousa, F.F., Rocha, M.G.C., Bargiela, P., Oliveira, A.C., Chem. Eng. J.,
168 (2011) 656–664.
38. de Sousa, H.S.A., de Assis F., Barros, A., Vasconcelos, S.J.S., Filho, J.M.,
Lima, C.L., Oliveira, A.C., Ayala, A.P., Junior, M.C., Oliveira, A.C.,
Appl. Catal. A: Gen., 406 (2011) 63–72.

b1469_Ch-13.indd 808 4/8/2013 12:41:52 PM


b1469 Catalysis by Ceria and Related Materials

Transformation of Oxygenated Compounds 809

39. Chai, S.-H., Wang, H.-P., Liang, Y., Xu, B.-Q., Green Chem., 9 (2007)
1130–1136.
40. Tae Kim, Y., Jung, K.-D., Duck Park, E., Bull. Korean Chem. Soc., 31
(2010) 3283–3289.
41. Suprun, W., Lutecki, M., Gläser, R., Papp, H., J. Mol. Catal. A: Chem.,
342–3 (2011) 91–100.
42. Yu, W., Zhao, J., Ma, H., Miao, H., Song, Q., Xu, J., Appl. Catal. A: Gen.,
383 (2010) 73–78.
43. Zheng, Y., Chen, X., Shen, Y., Chem. Rev., 108 (2008) 5253–5277.
44. Bancquart, S., Vanhove, C., Pouilloux, Y., Barrault, J., Appl. Catal. A:
Gen., 218 (2001) 1–11.
45. Barrault, J., Pouilloux, Y., Clacens, J.M., Vanhove, C., Bancquart, S.,
Catal. Today, 75 (2002) 177–181.
46. Ferretti, C.A., Soldano, A., Apesteguia, C.R., Di Cosimo, J.I., Chem.
Eng. J., 161 (2010) 346–354.
47. Abid, M., Touroude, R., Catal. Lett., 69 (2000) 139–144.
48. Abid, M., Ehret, G., Touroude, R., Appl. Catal. A: Gen., 217 (2001)
219–229.
49. Abid, M., Paul-Boncour, V., Touroude, R., Appl. Catal. A: Gen., 297
(2006) 48–59.
50. Sepúlveda-Escribano, A., Coloma, F., Rodríguez-Reinoso, F., J. Catal.,
178 (1998) 649–657.
51. Sepúlveda-Escribano, A., Silvestre-Albero, J., Coloma, F., Rodríguez-
Reinoso, F., Stud. Surf. Sci. Catal., 130 (2000) 1013–1018.
52. Silvestre-Albero, J., Rodríguez-Reinoso, F., Sepúlveda-Escribano, A.,
J. Catal., 210 (2002) 127–136.
53. Silvestre-Albero, J., Sepúlveda-Escribano, A., Rodríguez-Reinoso, F.,
Anderson, J.A., Phys. Chem. Chem. Phys., 5 (2003) 208–216.
54. Concepción, P., Corma, A., Silvestre-Albero, J., Franco, V., Chane-
Ching, J.Y., J. Am. Chem. Soc., 126 (2004) 5523–5532.
55. Silvestre-Albero, J., Coloma, F., Sepúlveda-Escribano, A., Rodríguez-
Reinoso, F., Appl. Catal. A: Gen., 304 (2006) 159–167.
56. Serrano-Ruiz, J.C., Luettich, J., Sepúlveda-Escribano, A., Rodríguez-
Reinoso, F., J. Catal., 241 (2006) 45–55.
57. Serrano-Ruiz, J.C., Huber, G.W., Sánchez-Castillo, M.A., Dumesic, J.A.,
Rodríguez-Reinoso, F., Sepúlveda-Escribano, A., J. Catal., 241 (2006)
378–388.

b1469_Ch-13.indd 809 4/8/2013 12:41:53 PM


b1469 Catalysis by Ceria and Related Materials

810 L. Vivier and D. Duprez

58. Concepción, P., Corma, A., Silvestre-Albero, J., Top. Catal., 46 (2007)
31–38.
59. Malathi, R., Viswanath, R.P., Appl. Catal. A: Gen., 208 (2001) 323–327.
60. Serrano-Ruiz, J.C., Sepúlveda-Escribano, A., Rodríguez-Reinoso,
F., Duprez, D., J. Mol. Catal. A: Chem., 268 (2007) 227–234.
61. Bachiller-Baeza, B., Rodríguez-Ramos, I., Guerrero-Ruiz, A., Appl.
Catal. A: Gen., 205 (2001) 227–237.
62. Campo, B., Volpe, M., Ivanova, S., Touroude, R., J. Catal., 242 (2006)
162–171.
63. Campo, B., Petit, C., Volpe, M.A., J. Catal., 254 (2008) 71–78.
64. Campo, B., Ivanova, S., Gigola, C., Petit, C., Volpe, M.A., Catal. Today,
133–5 (2008) 661–666.
65. Campo, B., Santori, G., Petit, C., Volpe, M., Appl. Catal. A: Gen., 359
(2009) 79–83.
66. He, L., Yu, F.-J., Lou, X.-B., Cao, Y., He, H.-Y., Fan, K.-N., Chem. Commun.,
46 (2010) 1553–1555.
67. Bhogeswararao, S., Srinivas, D., Catal. Lett., 140 (2010) 55–64.
68. Hasan, M.A., Zaki, M.I., Pasupulety, L., Appl. Catal. A: Gen., 243 (2003)
81–92.
69. Yamada, Y., Segawa, M., Sato, F., Kojima, T., Sato, S., J. Mol. Catal. A:
Chem., 346 (2011) 79–86.
70. Stubenrauch, J., Brosha, E., Vohs, J.M., Catal. Today, 28 (1996) 431–441.
71. Kunkes, E.L., Simonetti, D.A., West, R.M., Serrano-Ruiz, J.C., Gärtner,
C.A., Dumesic, J.A., Science, 322 (2008) 417–421.
72. Gaertner, C.A., Serrano-Ruiz, J.C., Braden, D.J., Dumesic, J.A., J. Catal.,
266 (2009) 71–78.
73. Gaertner, C.A., Serrano-Ruiz, J.C., Braden, D.J., Dumesic, J.A.,
ChemSusChem, 2 (2009) 1121–1124.
74. Gaertner, C.A., Serrano-Ruiz, J.C., Braden, D.J., Dumesic, J.A., Ind.
Eng. Chem. Res., 49 (2010) 6027–6033.
75. Serrano-Ruiz, J.C., Wang, D., Dumesic, J.A., Green Chem., 12 (2010)
574–577.
76. Serrano-Ruiz, J.C., Braden, D.J., West, R.M., Dumesic, J.A., Appl. Catal.
B: Environm., 100 (2010) 184–189.
77. Klimkiewicz, R., Tererycz, H., Grabowska, H., Morawski, I., Syper, L.,
Licnerski, B.W., J. Am. Oil Chem. Soc., 78 (2001) 533–535.

b1469_Ch-13.indd 810 4/8/2013 12:41:53 PM


b1469 Catalysis by Ceria and Related Materials

Transformation of Oxygenated Compounds 811

78. Klimkiewicz, R., Teterycza, H., React. Kinet. Catal. Lett., 75 (2002)
165–168.
79. Murkute, A.D., Jackson, J.E., Miller, D.J., J. Catal., 278 (2011)
189–199.
80. Lippert, S., Baumann, W., Thomke, K., J. Mol. Catal., 69 (1991)
199–214.
81. Idriss, H., Diagne, C., Hindermann, J.P., Kiennemann, A., Marteau,
M.A., J. Catal., 155 (1995) 219–237.
82. Raskó, J., Kiss, J., Appl. Catal. A: Gen., 287 (2005) 252–260.
83. Di Cosimo, J.I., Torres, G., Apesteguía, C.R., J. Catal., 208 (2002)
114–123.
84. Torres, G., Apesteguía, C.R., Di Cosimo, J.I., Appl. Catal. A: Gen., 317
(2007) 161–170.
85. Kunkes, E.L., Gürbüz, E.I., Dumesic, J.A., J. Catal., 266 (2009)
236–249.
86. Gürbüz, E.I., Kunkes, E.L., Dumesic, J.A., Appl. Catal. B: Environm., 94
(2010) 134–141.
87. Gangadharan, A., Shen, M., Sooknoi, T., Resasco, D.E., Mallinson,
R.G., Appl. Catal. A: Gen., 385 (2010) 80–91.
88. Skouta, R., Li, C.-J., Tetrahedron, 66 (2008) 4917–4938.
89. Corma, A., Garcia, H., Chem. Soc. Rev., 37 (2008) 2096–2126.
90. Casanova, O., Iborra, S., Corma, A., ChemSusChem, 2 (2009) 1138–1144.
91. Casanova, O., Iborra, S., Corma, A., J. Catal., 265 (2009) 109–116.
92. Smolentseva, E., Kusema, B.T., Beloshapkin, S., Estrada, M., Vargas, E.,
Murzin, D.Yu., Castillon, F., Fuentes, S., Simakov, A., Appl. Catal. A:
Gen., 392 (2011) 69–79.

b1469_Ch-13.indd 811 4/8/2013 12:41:53 PM


b1469 Catalysis by Ceria and Related Materials

This page intentionally left blank

b1469_Ch-13.indd 812 4/8/2013 12:41:53 PM


b1469 Catalysis by Ceria and Related Materials

CHAPTER 14

CERIA-BASED CATALYSTS FOR AIR


POLLUTION ABATEMENT
Anna Maria Venezia, Leonarda Francesca Liotta,
Giuseppe Pantaleo, and Alessandro Longo
Istituto per lo Studio dei Materiali Nanostrutturati (ISMN)-CNR,
Via Ugo la Malfa 153, 90146, Palermo, Italy

14.1 Introduction
Air pollution abatement is an important issue associated mainly with
the increasing demand for energy and with the volume production
of goods. A technological approach to the problem is environmental
catalysis, which provides appropriate materials for selectively and effi-
ciently converting noxious emission products to safe compounds.
Most of the catalytic processes for the removal of pollutants include
oxidation reactions with oxygen or air with low environmental impact.
In this respect ceria (CeO2) has been extensively used by virtue
of its unique electronic features.1 In contrast to the other rare-earth
elements (Ln), which normally exist as sesquioxides Ln2O3 with Ln3+
as their common valence state, cerium oxide is present mainly as
CeO2 due to the unstable [Xe] 4f 1 electronic configuration of Ce3+.
Ceria has a CaF2 fluorite structure consisting of a network of cerium
ions arranged in an fcc array.2 In certain temperature and oxygen
pressure conditions a range of intermediate oxides of composition
between Ce2O3 and CeO2 can exist.1 Reduced ceria arises with the
removal of O2− ions from the CeO2 lattice, generating anion vacancy
sites while maintaining the electrostatic balance by the reduction of
two cerium cations from the +4 to +3 oxidation state.3 Structural

813

b1469_Ch-14.indd 813 4/8/2013 12:42:32 PM


b1469 Catalysis by Ceria and Related Materials

814 A.M. Venezia et al.

investigation by x-ray diffraction (XRD) suggests that cubic Ce2O3


forms after the reduction of high-surface-area ceria and that
small ceria particles tend to stabilize the Ce3+ ions.3–4 Accordingly,
enhanced stability of surface oxygen vacancies and an enhanced
reducibility of small ceria particles compared to bulk ceria have
been observed. One important feature of CeO2 is its oxygen storage
capacity (OSC), which is the ability to adjust the oxygen content
when redox processes are involved. Moreover, when ceria is used as
a support, calorimetric measurements have demonstrated that the
adsorption energy of a metal such as Ag vapor to the (111) surface
of CeO2 is much higher compared to the adhesion energy to
another oxide like MgO (100), and it increases over thermally
reduced CeO1.8 (111).5 This result explains the sintering resistance
of late-transition-metal catalysts when supported over CeO2 and
therefore the suitability of ceria as a catalytic support.6
This chapter reviews recent advances in the use of ceria as a cata-
lyst component in environmentally important catalytic processes.
In particular, the oxidation of CO, the oxidation of volatile organic
compounds (VOCs), the combustion of methane, and the reaction
of NO with CO will be considered.

14.2 Oxidation of CO
Carbon monoxide along with unburned hydrocarbons is one of the
main toxic pollutants emitted from the exhaust of vehicle engines
and also from combustion of fossil fuels. One way to remove it is
through catalytic oxidation to CO2 in so-called “three-way catalysis”
(TWC), used in the after-treatment devices for automotive pollu-
tion control.7 Two classes of catalysts for CO oxidation are generally
considered: those based on noble metals (Pd, Pt, Rh, Au) and those
based on transition-metal oxides.7–11 In both series, ceria has often
been considered an important constituent. Indeed, as stated above,
the electronic configuration and the easy transformation between
the two cerium oxidation states, Ce4+ and Ce3+, determine its ability
to store and release oxygen. CeO2 is an “active oxide” in the sense
that, through its lattice oxygen, it participates in the oxidation reac-
tion, at variance with “inert” oxides like SiO2 or MgO, which act

b1469_Ch-14.indd 814 4/8/2013 12:42:32 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Catalysts for Air Pollution Abatement 815

only as carriers of the active species.12 With the inert class of oxides,
characterized by a low ability to adsorb oxygen, the reaction pro-
ceeds through the dissociative adsorption of oxygen on the sup-
ported metal particles. On active oxides like CeO2 molecular
oxygen can adsorb and then may or may not dissociate before
reacting with CO adsorbed on a supported noble metal or on the
same oxide system. Density functional theory corrected for on-site
Coulomb interactions (DFT+U) has been used to describe the oxy-
gen vacancies and the localization of the charge on Ce ions near to
vacancy sites.13–14 It has been ascertained that the oxygen vacancies,
after formation, make oxygen migration possible.13 It is generally
agreed that CO oxidation over CeO2 occurs by a Mars–van Krevelen-
type mechanism,15 involving the alternate reduction and oxidation
of surface ceria with the formation of surface oxygen vacancies
and their successive annihilation by gas-phase oxygen.16 Since sub-
traction of oxygen from the lattice requires a certain energy, the
mechanism on pure CeO2 occurs at a higher temperature. The low-
temperature CO oxidation activity of CeO2 can be promoted
through morphology and structure changes, by adopting an appro-
priate synthetic procedure, adding other oxides, or by supporting
noble metals.8,10,16–18 Examples from the recent literature, repre-
sentative of each of the different approaches for increasing oxygen
vacancy mobility and achieving higher catalytic efficiency, are
described in this section.

14.2.1 Nanostructured CeO2


As mentioned above, the catalytic activity of CeO2 in the CO oxida-
tion reaction can be enhanced by modifying the structural proper-
ties through specific syntheses. One of the major drawbacks of
ceria is its poor resistance towards sintering. Indeed, according to
a set of rules established several years ago, oxide sintering is favored
for metals having a high atomic weight like cerium.19 Moreover,
much effort has focused on increasing the OSC of ceria, when try-
ing to fabricate CeO2-based catalysts, by exposing reactive surfaces.
To this purpose, cerium oxides have been prepared using hexa-
decyltrimethylammonium bromide (CTAB) under hydrothermal

b1469_Ch-14.indd 815 4/8/2013 12:42:32 PM


b1469 Catalysis by Ceria and Related Materials

816 A.M. Venezia et al.

conditions.20 The synthesis involved the precipitation of cerium


hydroxide from the precursor Ce(NO3)3·6H2O with ammonia and
in the presence of CTAB, which acts as a structural template. The
reaction occurred in an autoclave at a fixed temperature and for a
certain length of time. By varying the molar ratio CTAB/Ce3+ and
by changing the reaction temperatures and the reaction times,
different ceria morphologies were obtained. For a CTAB/Ce3+ ratio
of 1.3 and for a reaction time of 24 h, nanoplates of CeO2 were
obtained. By increasing the reaction time, the nanoplates were
converted to nanorods. When the CTAB/Ce3+ ratio was 2.3 a quick
transformation from nanoplates to nanotubes occurred. In the
scheme in Fig. 14.1 the authors show the possible mechanism for
the formation of CeO2 with the different morphologies.20 At the
basic pH condition of the synthesis reaction CTA+ ions are adsorbed
on the CeO2 surface and preferentially on the (111) and (100)
planes, as demonstrated by Vantomme et al.21 A cubic plane struc-
ture forms, which, on increasing the temperature, becomes of
rhombic or hexagonal shape. Increasing the amount of CTAB with

Figure 14.1 Possible formation mechanism for CeO2 nanostructures. From Pan
et al.20 with permission.

b1469_Ch-14.indd 816 4/8/2013 12:42:32 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Catalysts for Air Pollution Abatement 817

respect to Ce3+ increases its capping ability therefore restricting


nanoplate growth and leading to its rolling and the formation of
nanorods and nanotubes. The structures obtained have been
tested in the CO oxidation reaction. The materials had a BET
surface area of 81 m2.g−1 for nanotubes, 52.5 m2.g−1 for nanorods,
and 37.2 m2.g−1 for nanoplates. A comparison of CO conversion as
a function of temperature for the different CeO2 morphologies is
shown in Fig. 14.2. In terms of T50, the temperature of 50% CO
conversion to CO2, nanoplates exhibited the lowest value (215 °C),
about 100 °C lower than reported previously for conventionally
prepared CeO2 powders. As described by Zhou et al.,17 CO oxida-
tion is strongly affected by the crystal planes due to their different
abilities for creating oxygen vacancies. On this basis, and in agree-
ment with transmission electron microscopy (TEM) analyses, a pos-
sible explanation for the better activity of the nanoplates is the
prevalent exposure of the (100) surfaces, which, according to com-
puter modeling,22 require less energy for oxygen vacancy formation
compared to the (111) and (110) surfaces. These results are in
agreement with an earlier study dealing with a series of CeO2 pow-
ders obtained by the calcination of samples of differently prepared
polycrystalline ceria.16 A careful high-resolution transmission elec-
tron microscopy (HRTEM) investigation was able to give accurate

Figure 14.2 Plots of CO conversion of various CeO2 nanostructures vs tempera-


tures. From Pan et al.20 with permission.

b1469_Ch-14.indd 817 4/8/2013 12:42:32 PM


b1469 Catalysis by Ceria and Related Materials

818 A.M. Venezia et al.

sizes for the crystallites, identifying also the particle facets. Overall
CO oxidation activity decreased with an increase of the sample
calcination temperatures, which increases the size of the ceria par-
ticles. When the specific activity per unit of surface area was calcu-
lated and plotted versus ceria crystallite size, the CO oxidation rate
can be seen to increase with size, regardless of the preparation
history of the samples. As shown by HRTEM analyses, calcination,
besides increasing the particle size with a consequent decrease of
the surface area, contributes to restructuring the particles with
exposure of the more reactive (100) planes.
Another method recently reported for the preparation of cata-
lytically active CeO2 material is the use of the low-cost biopolymer
chitosan as a template.18 The synthesis consisted of adding together
chitosan dissolved in acetic acid and aqueous cerium nitrate in a
variable weight ratio. After precipitation with ammonia followed by
drying and calcining at the desired temperature, the cerium oxides
obtained were characterized by high surface area (∼144 m2.g−1). The
area changed as a function of the calcination temperature and also
as a function of the weight ratio Ce3+/chitosan. Increasing the tem-
perature and increasing the amount of biopolymer had the opposite
effect. TEM images showed sheets of mesoporous aggregates of
cubic CeO2 nanocrystals. The materials evaluated in the CO oxida-
tion reaction were quite active with the best catalyst giving 90%
CO conversion at 320 °C (gas hourly space velocity or GHSV of
60000 h−1), a much improved result compared to commercial CeO2.
The improvement in the activity was correlated with an increase of
the lattice parameters and the lattice volume due to chitosan-assisted
synthesis.18 In accord with Ho et al. the expansion of the lattice vol-
ume of ceria, favoring the replacement of Ce4+ of radius 0.97Å, with
Ce3+ of radius 1.14 Å, increases the redox properties of ceria.23

14.2.2 Ceria-based oxide catalysts


14.2.2.1 CoOx/CeO2 system
Of the ceria-based oxide catalysts, mixed ceria cobalt and mixed
ceria–copper have been the most extensively investigated.9–10,24

b1469_Ch-14.indd 818 4/8/2013 12:42:33 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Catalysts for Air Pollution Abatement 819

Composite catalysts, CoOx/CeO2, with a Co3O4 weight composition


of 5, 10, 15, and 20%, have been prepared by co-precipitation start-
ing from an aqueous solution of cerium and cobalt nitrates and
adding sodium carbonate up to pH 8.24 After filtering and washing,
the precipitate was calcined in air at 650 °C for 5 h. CO oxidation
catalytic tests were performed under stoichiometric conditions.
A comparison of composites of ceria with other transition-metal
oxides, like CuO, MnO2, NiO, and Cr2O3, similarly prepared,
showed the superiority of the CoOx/CeO2 system in spite of its lower
surface area. The optimum cobalt oxide content was 15 wt%. On
this sample, the effect of calcination temperature on the light-off
temperature for CO oxidation was investigated. Table 14.1 summa-
rizes the variation of T50 with structural and morphology data.
Besides the effect on the surface area and particle size, the calcina-
tion temperature also influenced the surface cobalt concentration,
with a maximum for the sample treated in air at 650 °C, which was
also the most active catalyst. In accord with XRD results showing
separate CeO2 and Co3O4 phases, the increase of the cobalt surface
concentration was an indication that cobalt oxide was not soluble in
CeO2 and that a balance between the surface segregation of cobalt
oxide and its crystallite size gives the optimum catalytic performance
of the 650 °C treated sample. X-ray photoelectron spectroscopy
(XPS) analyses of the CoOx/CeO2 system revealed the effect of the
ceria on the mobility of the oxygen on the cobalt oxide. Based on
the change of the intensity ratio between the satellite shake-up peak,

Table 14.1 T50, BET surface area, Co/Ce atomic ratios derived from XPS and
Co3O4 particle sizes derived from XRD for CoOx/CeO2 (15%) after 5 h in air at
different calcination temperatures. (Adapted from Kang et al.24)

Tcalc (°C) T50 (°C) S(BET) (m2.g−1) Co/Ce dCo3O4 (Å)


450 165 56 0.8 n.d.
550 145 55 1.0 140
650 125 44 2.0 160
750 135 38 1.7 200
850 149 33 1.3 210

b1469_Ch-14.indd 819 4/8/2013 12:42:33 PM


b1469 Catalysis by Ceria and Related Materials

820 A.M. Venezia et al.

characteristic of Co2+ species, and the main Co 2p3/2 peak, it was pos-
sible to ascertain the enhancement of the Co3+ formation in the
presence of ceria.
Luo et al.25 prepared a series of composite Co3O4–CeO2 catalysts
using a surfactant template (ST) method and compared them with
samples prepared by the impregnation (IM), co-precipitation (CP),
or combustion (CB) methods. The ST procedure consisted in pre-
cipitating the cerium and cobalt hydroxides in the presence of a
surfactant, CTAB, followed by thermal aging at 120 °C and calcina-
tion at 500 °C. According to nitrogen adsorption-desorption iso-
therms, mesoporous materials with uniform pore sizes (ranging from
36 Å to 61 Å with increasing cobalt loading) and quite high surface
area (∼100 m2.g−1) were obtained. The sample CeCo30 from the ST
procedure, with atomic ratio Co/(Co+Ce) = 0.30, tested in the oxida-
tion of CO (weight hourly space velocity or WHSV = 10000 ml.g−1.h−1),
exhibited the best activity with T50 = 94 °C. This value was approxi-
mately more than 30 °C below those obtained with homologous
samples prepared by different procedures. The redox properties,
revealed by the temperature-programmed reduction by H2 (H2-TPR)
patterns shown in Fig. 14.3, were very much dependent on the type
of synthesis. Pure CeO2 pattern exhibited a peak at 820 °C attributed
to bulk CeO2 and visible in all composite samples and a peak at
528 °C, due to surface oxygen species, overlapping with Co reduction
peaks. The reduction of Co3O4 depended strongly on particle size.
According to Spadaro et al.26, large particles not interacting with
other oxides are reduced directly to metallic cobalt in a temperature
range between 320 °C and 480 °C. Small Co3O4 particles interacting
with ceria are reduced in two steps, the first characterized by a low-
temperature peak at around 240–320 °C, due to the reduction of
Co3+ species to Co2+ at the interface with ceria, and a high-tempera-
ture peak in the range of 480–700 °C due to the reduction of Co2+,
interacting with CeO2, to metallic Co.25 Since the peaks in the range
320–480 °C are due to the independent or bulk-like Co3O4 phase,
their intensity gives an indication of the dispersion of the Co3O4
phase. As shown in Fig. 14.3, according to the intensity of this peak,
the dispersion of Co3O4 in the surfactant-template-prepared sample

b1469_Ch-14.indd 820 4/8/2013 12:42:33 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Catalysts for Air Pollution Abatement 821

Figure 14.3 H2-TPR profiles of (a) CeCo20 prepared by different methods, Co3O4
and CeO2; and (b) catalysts prepared by the surfactant template method. From Luo
et al.25 with permission.

b1469_Ch-14.indd 821 4/8/2013 12:42:33 PM


b1469 Catalysis by Ceria and Related Materials

822 A.M. Venezia et al.

(CeCo20) was the largest with respect to the other samples. A further
promotion of the activity of the composite system CeCo30 was
achieved by depositing 0.5 wt% of palladium through impregnation.
As shown in Fig. 14.3(b), the presence of palladium enhanced the
reduction of the cobalt species associated with the low-temperature
shift of the two cobalt peaks at 154 °C and at 368 °C. A diffuse reflec-
tance IR Fourier-transform spectroscopy (DRIFTS) study of the
CeCo30 and Pd/CeCo30 samples, after exposure to a CO + O2 gas
mixture at 25 °C, revealed bands attributable to bidentate carbonate,
followed by the appearance of bands due to gas-phase CO2. These
bands suggested that oxidation of CO at the ambient temperature
occurred through the formation of surface carbonates as intermedi-
ates. For the Pd-promoted samples, similar bands but of lower inten-
sity were observed. Moreover, in this case, there was no direct relation
between the observed CO and CO2 bands and the activity. To account
for these results, different reaction mechanisms over the unpro-
moted and palladium-promoted CeCo30 samples were suggested as
shown in Fig. 14.4. According to the proposed mechanism, the loss
of activity observed for the unpromoted mixed-oxide catalyst was
related to the regeneration of oxygen vacancies, which was assumed
to be the rate-determining step needed for oxygen activation. For the
palladium-promoted catalyst, as shown in Fig. 14.4(b), Pd species,
like other noble metals, are able to activate oxygen by dissociating it
even at low temperatures. The atomic oxygen then spilled over the
basic support reacting with the adsorbed carbonate. The synergism
between Pd, able to activate oxygen, and the ST-prepared CeCo30
oxide, able to adsorb CO, produced a catalyst with better perfor-
mance for low-temperature CO oxidation compared to Pd supported
over ceria or over Co3O4.

14.2.2.2 CuO/CeO2
Another interesting composite system based on ceria are CuO/CeO2
mixed oxides. This system has been widely studied as a replacement
for the more expensive noble-metal systems and, with respect to the
cobalt system, it has the advantage of being environmentally greener

b1469_Ch-14.indd 822 4/8/2013 12:42:33 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Catalysts for Air Pollution Abatement 823

Figure 14.4 Proposed CO reaction mechanism over (a) CeCo30 and (b) Pd/
CeCo30. From Luo et al.25 with permission.

and also more tolerant to SO2.11 Several different methods of prepa-


ration have been tried, from inert-gas condensation (IGC), impreg-
nation, sol-gel and thermal synthesis coupled with impregnation

b1469_Ch-14.indd 823 4/8/2013 12:42:33 PM


b1469 Catalysis by Ceria and Related Materials

824 A.M. Venezia et al.

methods.27–29 As in the Co3O4/CeO2 system, the main factors control-


ling the activity in low-temperature CO oxidation are the nanostruc-
tured morphology and the dispersion of copper at the sample
surface. Substantial differences in CO oxidation activity and in struc-
tural properties were observed between a series of CuO/CeO2 sam-
ples (copper loading ranging from 1.3 wt% to 15.1 wt%) prepared
by copper impregnation of a sol-gel prepared CeO2 (surface area
∼ 68 m2.g−1) and a series of analogous catalysts prepared by impregna-
tion of a commercial ceria of much lower surface area (∼ 4 m2.g−1).28
The samples over the sol-gel ceria were much more active compared
to those over commercial ceria. Of all the different preparation pro-
cedures, the one based on wet impregnation of CeO2, obtained by
thermal decomposition of Ce(NO3)2·6H2O with an aqueous solution
of copper nitrate produced the best catalysts.29 The thermal treat-
ment to produce the CeO2 and the catalyst calcination temperature
had a strong influence on activity. In particular, it was found that the
best catalyst was obtained over CeO2 (600 °C), which was a ceria
derived from thermal decomposition carried out at 600 °C. CO con-
version rates for the CuO/CeO2 (600 °C) samples calcined at differ-
ent temperatures are shown in Fig. 14.5. It is important to note that
the activity of the pure CuO oxide was almost zero under the same
reaction conditions.29 As with cobalt, the origin of the different levels
of catalytic activity relies on the redox properties of the materials.
H2-TPR patterns of the sample CuO/CeO2 (600 °C) calcined at
300 °C and 800 °C were quite different. The patterns for the sample
calcined at 300 °C had a strong peak (β) at about 199 °C with a shoul-
der at 218 °C, associated with the reduction of larger CuO particles
weakly interacting with ceria, and a small peak at 140 °C attributed
to the reduction of non-crystalline CuO strongly interacting with
CeO2. The patterns for the sample calcined at 800 °C showed a main
peak at 248 °C (γ) attributed to the reduction of larger CuO particles
not associated at all with ceria. XRD data confirmed the presence of
a CuO crystal phase along with the solid solution Cu–Ce–O. Indeed,
as evidenced by the decrease of the lattice parameter from 5.409 Å
for pure CeO2 to 5.402 Å for the CuO/CeO2 system, a certain degree
of copper incorporation into the CeO2 lattice occurred, as the Cu2+

b1469_Ch-14.indd 824 4/8/2013 12:42:33 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Catalysts for Air Pollution Abatement 825

Figure 14.5 Catalytic activity of CeO2–C (from thermal decomposition of cerous


nitrate at 600 °C) and supported CuO catalysts calcined at different temperatures.
From Zheng et al.29 with permission.

ion radius (0.076 Å) is smaller than that of Ce4+ (1.010 Å). XPS data
suggested that the oxidation state of copper in the mixed-oxide sam-
ples was lower than Cu2+. To conclude, once again the role of ceria
in these types of system appeared to be a modification of the chemi-
cal and structural properties of the second oxide by enhancing the
redox capability.
Because of the excellent catalytic performance of the ceria-
based Co3O4 and CuO catalysts, a combination of the two systems
was investigated by preparing, through co-precipitation, mixed Cu–
Co–Ce–O oxides with different compositions.20 The mixed oxides
had a larger surface area with respect to the Co–Ce–O and Cu–
Ce–O composite oxides. Moreover, the addition of cobalt to the
Cu–Ce–O catalyst enhanced thermal stability because the particle
size was unchanged on calcination at 850 °C. According to the
H2-TPR patterns, the peaks due to cobalt oxide reduction and those
for copper oxide reduction were reciprocally affected and shifted
towards lower temperatures. The interaction between cobalt and

b1469_Ch-14.indd 825 4/8/2013 12:42:33 PM


b1469 Catalysis by Ceria and Related Materials

826 A.M. Venezia et al.

copper enhanced sample reducibility, weakening the surface Cu–O


and Co–O bonds and therefore favoring oxygen transfer and pro-
moting the low-temperature CO oxidation activity of the catalyst.

14.2.3 Ceria-supported noble-metal catalysts


As described above, by using an appropriate synthetic procedure or
adding other oxides to CeO2, it is possible to enhance CO oxidation
activity of ceria-based catalysts, achieving under the most favorable
conditions a T50 of about 90 °C. Nevertheless, a drawback of using
base-metal oxides is their sensitivity to deactivation by sulfur and
water.11 For low-temperature oxidation, as required in sensors or air
depollution treatments, precious metals are still necessary. The
development of catalytic converters, in the 1970s in the United
States and in the 1990s in Europe, highlighted the exceptional
activity of noble metals (Pt, Pd, Rh) for CO oxidation and the
specific role of oxides with a high oxygen storage capacity (OSC),
which allow the metal to continue working in cyclic conditions.11,30
Of the noble metals, palladium exhibits the highest activity for CO
oxidation.7,30 As already mentioned for Pd-promoted CeCoO cata-
lysts, the reaction over a noble metal supported on ceria involves a
cooperative effect between the metals and the oxide, the so-called
dual site mechanism.11 A recent study by temporal analysis of prod-
ucts (TAP) experiments for Pt supported over ceria showed two
independent sites with different activities.31 The high activity site was
associated with the metal/support interface and the low activity one
was located on the support. The different sites were characterized by
two different activation energies. Moreover, at variance with the
stable number of high activity sites, the number of the low active
sites increased with reaction temperature.
The size of the metal particle plays an important role in the
structure-sensitive reaction when site coordination (edge, kink, or
terrace sites) or crystal orientation affect the activity and selectivity.32
In recent work, transient curves of gas-phase CO2 produced with
CO-pulse injection on Pd/CeO2 at 500–700 °C, established the
structure sensitivity of the CO oxidation reaction on supported

b1469_Ch-14.indd 826 4/8/2013 12:42:33 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Catalysts for Air Pollution Abatement 827

palladium catalysts. The experiment established that the oxygen


storage capacity range was strongly dependent on Pd particle
size.33,34 A mathematical model based on a Eley–Rideal mechanism
was developed to describe the dynamic response of CO during the
pulse experiment. The study indicated that CO interacted with the
oxygen of the pre-oxidized palladium, with the vacant sites created
on the PdO being replenished by a back-spillover mechanism from
the ceria, which acts as an oxygen pool.34 The simulations indicated
that both the transient rate of CO oxidation and that of oxygen
spillover decreased with increasing Pd particle size. Indeed when
the metal particles were small the interfacial contact between ceria
and Pd was large and therefore the rate of oxygen back-spillover
increased. As a consequence, the method of synthesis and the cata-
lyst treatment are crucial for CO oxidation activity.35 In particular, it
was shown for Pd and Pt catalysts supported over CeO2 that the syn-
thesis by combustion, involving an increase at a high temperature
(1000 °C) for short periods, favored the Pd–Ce interaction. It was
suggested that this interaction would lead to the formation of a solid
solution between the metal and the CeO2 with the creation of oxy-
gen vacancies and ceria in a reduced state (Ce3+).36
Gold-based catalysts have been extensively studied in the last two
decades, since the surprisingly high activity in CO oxidation at low
temperatures was reported for small (30–50 Å) gold particles sup-
ported on various oxides.37 In particular, the combination of gold
with ceria attained the lowest temperature for CO oxidation, well
below the temperature achieved with other noble metals.8 Again, as
discussed above for palladium and platinum catalysts, besides the size
of the metal particles, the interaction between the gold and the sup-
port is important. A debatable issue is the nature of the active sites,
whether they are metallic Au or charged Au+ and Au3+ species.37,38
In the former, size would play the main role, whereas in the latter the
interaction with the support would be more important.
In order to discriminate between the activity of the different
gold species on ceria, our group prepared ceria-supported gold cata-
lysts by classical deposition-precipitation (DP) followed by drying at
393 K, and by the solvated metal atom dispersion (SMAD) technique

b1469_Ch-14.indd 827 4/8/2013 12:42:33 PM


b1469 Catalysis by Ceria and Related Materials

828 A.M. Venezia et al.

followed by calcination at 623 K to remove the residual organic


solvent.8 For comparison, a catalyst prepared by co-precipitation
(CP) and calcined at 673 K was also used. The samples, character-
ized by XRD, TPR, and XPS, were tested in the low-temperature
oxidation of CO. In Fig. 14.6, CO conversion as a function of tem-
perature from 173 K up to 723 K is displayed for pure ceria and for
the differently prepared samples. The addition of gold substantially
improved CO oxidation activity of the pure support; however, its
effect strongly depended on the preparation method, with the larg-
est activity exhibited by the DP sample, followed by the SMAD sam-
ple, and with the CP sample showing the lowest activity. The DP
sample dried at 393 K was exceptionally active below room tempera-
ture with a specific rate of 6×10−6 molCO.sec−1.gcat−1 at 263 K and yield-
ing total CO conversion at room temperature. The rate was higher
than the value 2.2×10−6 molCO.sec−1.gcat−1 reported for gold over
nanocrystalline CeO239 and accordingly the sample appeared to be
the most active catalyst so far reported for this type of reaction.
As shown by the TPR patterns, the presence of gold caused a

Figure 14.6 CO conversion percentage as a function of temperature for different


catalysts. From Venezia et al.8 with permission.

b1469_Ch-14.indd 828 4/8/2013 12:42:33 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Catalysts for Air Pollution Abatement 829

low-temperature shift of the ceria surface reduction peak.8,40 The


shift depended on the preparation procedure following a trend simi-
lar to CO conversion. Furthermore the Au 4f photoelectron spectra,
shown in Fig. 14.7, for the sample obtained by DP and the sample
obtained by SMAD, clearly show the different oxidation states of
gold. According to the binding energy of the main peaks, AuCe(DP)
contained ionic Au+ and Au3+, whereas the SMAD sample contained
mainly metallic Au with a very small amount of Au+.41,42 A small shift
of the CeO2 lattice parameter was observed in the XRD patterns of
the Au/CeO2(DP) sample only. This suggested a certain degree of
gold incorporation into the ceria fluorite structure, which prevented
gold particle sintering and made this particular catalyst thermally
stable. In summary, the remarkable behavior of the sample prepared
by the DP procedure was attributed to the presence of ionic gold in
intimate contact with ceria, contributing to Ce–O weakening and to

Figure 14.7 Experimental and fitted Au 4f photoelectron peaks of the samples


prepared by deposition-precipitation and by SMAD. The solid and dashed lines
indicate two different doublet components. From Venezia et al.8 with permission.

b1469_Ch-14.indd 829 4/8/2013 12:42:33 PM


b1469 Catalysis by Ceria and Related Materials

830 A.M. Venezia et al.

an increase in oxygen mobility. The presence of ionic gold, mainly


Au+1 species, detected by XPS analyses, seemed the main requisite
for the highest CO conversion at the lowest temperature.41,42

14.3 Oxidation of VOCs


Volatile organic compounds (VOCs) are major contributors of air
pollution. They act as toxic compounds and as precursors of ozone
and photochemical smog. They are produced by outdoor sources
like industrial processes and automobile exhausts and also by indoor
sources, like household cleaning products and office supplies.43,44
VOCs, characterized by boiling points in the range 50–260 °C at
atmospheric pressure and low water solubility, include a variety of
chemicals like benzene, formaldehydes, toluene, propene phenol,
acetone, styrene, and halogenated hydrocarbons. The most efficient
methods for removing VOCs are thermal and catalytic combustion.
The latter method is becoming more attractive because of the much
lower operating temperature compared to thermal burning, and
because no undesirable compounds such as dioxins and nitrogen
oxides are produced. Moreover the method can be used with a
lower concentration of pollutants (less than 1000 ppm). There is
much ongoing effort in the development of catalytic materials with
high performance in terms of both activity and stability. Noble-
metal-based catalysts such as platinum, palladium, and rhodium,
generally supported on SiO2 or Al2O3, are quite active for the low-
temperature complete oxidation of VOCs. The catalytic perfor-
mance of the different noble metals varies with the nature of the
VOCs. Moreover, the presence of CO influences the activity of the
catalysts due to the competition between the adsorption of CO and
organic compounds on the metal surface.45 Due to the high costs
and sensitivity to chlorine poisoning, researchers have been trying
to replace noble-metal catalysts with metal-oxide-based catalysts,
which are less active but more resistant to poisoning.44 The following
sections give an overview of the most recent research and the
progress over the past 10 years on ceria-based oxides and ceria-
supported noble-metal systems.

b1469_Ch-14.indd 830 4/8/2013 12:42:33 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Catalysts for Air Pollution Abatement 831

14.3.1 Ceria and ceria-based mixed oxides


CeO2 and CeO2-based mixed oxides are cheap, environmentally
friendly and efficient catalysts for the catalytic decomposition of
VOCs.44 As with the oxidation of CO, VOC oxidation reactions over
ceria are considered to follow a Mars–van Krevelen mechanism in
which the key steps are the supply of oxygen by the readily reducible
oxide and its re-oxidation by oxygen.46–48 A very active CeO2 catalyst
prepared by thermal decomposition of cerium nitrate, followed by
calcination at 550 °C, was reported for the catalytic conversion of
trichloroethylene (TCE), which is used as a model for chlorinated
volatile organic compounds (CVOCs).49 The authors studied the
effect of variables, like the calcination temperature and the presence
of water vapor, on catalytic activity. The selectivity to HCl or Cl2 with
or without water was investigated. A characterization by XRD revealed
a fluorite structure of the CeO2 with crystallite particle size increasing
from 130 Å to 360 Å with calcination temperature.49 The presence of
basic sites in terms of strength and numbers was investigated by CO2-
TPD (temperature-programmed desorption of CO2). The basicity
was correlated with the number of oxygen defects produced at differ-
ent calcination temperatures. Quite interestingly, the sample cal-
cined at the lowest temperature of 450 °C and the one calcined at the
highest temperature of 800 °C were less basic. Indeed both had fewer
oxygen defects, the former being more stoichiometric and the latter
having the largest CeO2 crystallite size. A compromise between the
two was observed with the sample calcined at 550 °C, which had the
highest number of basic sites. The conversion rates of TCE on CeO2
catalysts calcined at different temperatures as a function of reaction
temperature is shown in Fig. 14.8. The catalyst calcined at 550 °C
exhibited the highest activity with 90% conversion of trichloroethyl-
ene at 205 °C. This performance was attributed to the smaller particle
sizes, larger surface area, and superior basicity. However, the catalyst
was not thermally stable and deactivated in a few hours due to the
strong adsorption of HCl and Cl2. No coke and no chlorination of
the ceria supports were observed. From XPS investigation of the
fresh and aged catalysts it was concluded that deactivation was due to

b1469_Ch-14.indd 831 4/8/2013 12:42:33 PM


b1469 Catalysis by Ceria and Related Materials

832 A.M. Venezia et al.

Figure 14.8 TCE combustion light-off curves over CeO2 catalysts calcined at differ-
ent temperatures: „ 550 °C; … 450 °C; z 650 °C; { 800 °C. Gas composition: 1000
ppm TCE, air balance; GHSV = 15000 h−1. From Dai et al.49 with permission.

the adsorption of electro-negative chlorine species neutralizing the


basic sites of ceria, which are important for the catalytic combustion
of TCE, and also inhibiting the regeneration of the surface oxygen,
which is consumed during catalytic oxidation.49
The influence of preparation conditions of nanocrystalline ceria
catalysts on the total oxidation of naphthalene was recently investi-
gated.50 The ceria was prepared by homogeneous precipitation with
urea. The gel obtained was kept under reflux for a specific time and
then dried and calcined. Different samples were obtained at differ-
ent calcination temperatures and times, and different aging times
corresponding to the period spent by the catalyst under reflux dur-
ing preparation. Varying these conditions resulted in differences in
surface area, crystallite size, oxygen defect concentration (as shown
by the full width at half maximum (FWHM) of the characteristic
Raman band), morphology (as seen from scanning electron micros-
copy images), and reducibility of the CeO2 samples. The best catalyst
had 100% conversion of CO2 at 200 °C with a gas hourly space veloc-
ity equal to 25000 h−1.
Modification of ceria by the addition of a second oxide has
been discussed as a way of increasing the thermal stability and

b1469_Ch-14.indd 832 4/8/2013 12:42:33 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Catalysts for Air Pollution Abatement 833

modifying the electronic properties, which ultimately affect oxygen


mobility. Gonzalez-Velasco and co-workers investigated the com-
plete oxidation of VOCs and CVOCs over ceria–zirconia systems.50,51
They used a series of mixed oxides (CexZr1-xO2) with varying molar
composition prepared by precipitation from nitrate precursors fol-
lowed by calcination at 550 °C. Samples with a cerium loading
above 68% had a cubic structure whereas samples with lower ceria
content had a tetragonal structure. Moreover, acidity in terms of
strength and number of acid sites, determined from NH3-TPD
measurements, was found to increase with zirconia content.50 IR
investigation confirmed the presence of only Lewis acidity. H2-TPR
patterns of the mixed oxides presented mainly one peak between
525 and 550 °C, whereas pure CeO2 had a low-temperature peak at
500 °C, which was assigned to the reduction of the surface, and a
peak at 900 °C assigned to the reduction of the bulk. The partial
replacement of Ce4+ with Zr4+ in the CeO2 lattice leads to a distor-
tion of the cubic lattice resulting in higher oxygen mobility, allow-
ing easy reduction of the bulk.52 The mixed-oxide samples were
tested in the combustion of a binary mixture of 1,2-dichloroethane
(DCE) or trichloroethylene (TCE) with n-hexane (HEX) to mimic
the feedstream found in industrial applications.51 The catalytic
results are summarized in Table 14.2 in terms of T50 and T90 for the
decomposition of each VOC molecule in the corresponding binary
mixture. It is clear that the incorporation of zirconia into the ceria
lattice decreases the light-off temperature of the chlorinated com-
pounds of which the saturated compounds were more easily oxi-
dized.51 The promoting effect of zirconia was associated with
surface acidity, which depended on the composition of the mixed
oxide.50 Accordingly, it was suggested that the first step of the com-
bustion process was the adsorption of a chlorinated molecule on a
strong acid site, which was then attacked by mobile oxygen species
from the solid solution through a Mars–Van Krevelen mechanism.
As with pure CeO2 catalysts49 no coke formation occurred, probably
due to the absence of Brønsted acidity as shown by an IR investiga-
tion. The conversion of n-hexane in the mixture in comparison
with the conversion of the hexane singly fed was inhibited to a large

b1469_Ch-14.indd 833 4/8/2013 12:42:34 PM


b1469 Catalysis by Ceria and Related Materials

834 A.M. Venezia et al.

Table 14.2 T50 ( °C) and T90 ( °C) for the decomposition of each VOC in the
corresponding binary mixture over CexZr1-xO2 catalysts. From Gutierrez-Ortiz et al.51
with permission.

Catalyst HEX DCE TCE


With DCE With TCE With TCE With HEX
CeO2 T50 235 260 330 450
T90 295 330 400 540
Ce0.8Zr0.2O2 T50 270 270 325 435
T90 335 350 395 540
Ce0.68Zr0.32O2 T50 275 270 320 435
T90 360 370 390 530
Ce0.5Zr0.5O2 T50 280 285 315 440
T90 390 380 380 535
Ce0.15Zr0.85O2 T50 294 295 325 450
T90 430 425 390 530
ZrO2 T50 305 305 365 490
T90 490 495 410 550

extent when unsaturated TCE was present. Likewise the conversion


of TCE and DCE were inhibited when n-hexane was present in the
feed. This inhibition, more marked with the TCE, was explained in
terms of the competitive adsorption of the VOCs on the catalyst
surface.53 It should be pointed out that the presence of HEX also
influenced the product distribution; principally it increased the
amount of HCl with respect to Cl2. The effect was attributed to the
formation of water in the product stream enhanced by HEX acting
as a hydrogen-supplying compound.51 The water may indeed dis-
sociate into H+ and OH− preferentially forming HCl or may shift the
Deacon reaction (2HCl + 1/2O2 = Cl2 + H2O ) to the left with the
production of HCl.
In general the production of Cl2 during the combustion of chlo-
rinated compounds depends on the type of feed, particularly on the
hydrogen content. For a hydrogen-rich molecule like dichloroeth-
ane, it was observed that ceria alone or ceria-enriched samples pro-
duced more Cl2 compared to zirconia-enriched samples. Unlike a

b1469_Ch-14.indd 834 4/8/2013 12:42:34 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Catalysts for Air Pollution Abatement 835

chlorinated VOC, the conversion of n-hexane was not promoted by


the addition of zirconia but ceria alone exhibited the lowest com-
bustion temperature and the highest yield of CO2. Conversely pure
zirconia, which is a poor oxidation catalyst, produced a large
amount of CO. On the basis of the catalytic oxidation of the three
different molecules it was concluded that oxidation of chlorinated
compounds was more demanding and required active oxygen spe-
cies from the bulk of the catalyst, whereas the combustion of n-hex-
ane was primarily controlled by surface oxygen species.51
MnOx–CeO2 and CuO–CeO2 mixed oxides have recently been
investigated for the catalytic oxidation of VOCs.54,55 A series of
mixed MnOx–CeO2 oxides with Mn/(Mn+Ce) = 5 were prepared
with three different methods, co-precipitation, modified co-precipi-
tation, and sol-gel, and tested in the catalytic oxidation of formalde-
hyde, which is regarded as a major indoor pollutant emitted from
building and decorative materials.56 Briefly, the sol-gel method con-
sisted in adding citric acid to a solution of Mn(NO3)2 6H2O and
(NH4)2Ce(NO3)6, heating to 323 K, and stirring until the formation
of yellowish gel, which was finally calcined at 773 K (sample SG-773).
The co-precipitation method consisted in precipitating, using KOH,
hydroxides from transition-metal nitrate precursors, followed by
aging, filtration, and calcination at 773 K (sample CP-773). The
third method was modified co-precipitation in which KMnO4 was
also added to the other precursors (sample MP-773). The conver-
sion rate of HCHO as a function of temperature is given in Fig. 14.9
for the three mixed-oxide samples. The crucial influence played by
the preparation method is clear, with the sample MP-773 being the
most active. The very low activity for CeO2 on its own, MnOx on its
own, and their physical mixture, in the same temperature range,
confirmed that there is a strong synergism in the mixed oxides. The
surface areas and average pore diameters were not responsible for
the difference in activity since very similar values were observed for
MP-773 and CP-773. Quite different H2-TPR patterns were obtained
for the mixed and single oxides.56 Pure CeO2 exhibited a broad peak
between 630 K and 890 K due to the removal of surface oxygen.
Pure MnOx showed two overlapping peaks at 623–843 K with a small

b1469_Ch-14.indd 835 4/8/2013 12:42:34 PM


b1469 Catalysis by Ceria and Related Materials

836 A.M. Venezia et al.

Figure 14.9 Temperature dependence of HCHO conversion over MnOx–CeO2


catalysts: HCHO = 580 ppm, O2 = 18.0%, He balance, GHSV = 21000 mL.gcat−1.h.
From Tang et al.56 with permission.

shoulder at 590 K. From the total amount of hydrogen used it was


considered that the initial state of manganese in MnOx was Mn4+
with a final reduction to MnO.57 In comparison with pure oxides,
the reduction temperatures of the mixed samples were shifted to
lower temperatures and exhibited different shapes depending on
the preparation procedure. This effect was attributed to a mutual
interaction between the two oxides. The TPR patterns of sample
MP-773 showed two intense peaks, one at 644 K attributed to the
reduction of MnOx to Mn3O4 and another at 703 K attributed to the
simultaneous reduction of Mn3O4 to MnO and the reduction of sur-
face ceria. The position of the peaks along with larger hydrogen
consumption pointed to the presence of Mn4+ species in the MP-773
sample and the presence of Mn3+ in the other two mixed oxides.
The higher oxidation state of manganese in this sample was also

b1469_Ch-14.indd 836 4/8/2013 12:42:34 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Catalysts for Air Pollution Abatement 837

supported by XPS results. According to XRD analyses, the mixed


samples contain only the cubic CeO2 phase. However, only the lat-
tice parameter for the MP-773 sample, slightly smaller than the lat-
tice parameter of pure ceria (0.541 nm), was indicative of a solid
solution between MnOx and CeO2. A decrease of the lattice con-
stant of ceria in similar systems was evidenced also by Kang et al.,
who established a solubility limit of 5–10 at% for Mn ions in CeO2.58
On the basis of these structural properties, the formation of the
MnOx–CeO2 solid solution appears to be crucial for the complete
oxidation of formaldehyde at lower temperatures. The superior
activity of sample MP-773 was explained by the synergistic interac-
tion between manganese and cerium oxides in which the MnO2
species released active oxygen, which participated in the oxidation
of HCHO, and then through oxygen transfer from CeO2, the Mn2O3
was oxidized again to MnO2.
Other manganese and cerium mixed oxides, denoted as MnxCe1-x
with x=Mn/(Mn+Ce)= 0.05, 0.15, 0.25, 0.50, and 0.75 (as atomic
ratios), were prepared by the combustion method and utilized in
the total oxidation of a gaseous feed composed of ethanol, ethyl
acetate, and toluene, chosen as representative VOCs.54 The combus-
tion method involved the use of urea as a fuel and nitrates as precur-
sors. The nitrate salts of manganese and cerium were mixed with an
appropriate amount of urea, heated to 80 °C, and then the viscous
gel was placed into a tubular furnace at 500 °C. After a short time,
ignition took place with a rapid evolution of gases yielding a volumi-
nous powder. The powder was calcined at 550 °C to remove residual
carbon components. The specific surface areas of the samples
increased from about 10 m2.g−1 for pure MnOx to a maximum of 59
m2.g−1 for a sample containing 95 at% Ce, and back to about 5 m2.g−1
for pure CeO2. The absence of manganese-oxide phases in the
Ce-rich samples was attributed to the formation of a solid solution.
Indeed, from a consistent decrease of the ceria lattice parameter
with increasing manganese content, the incorporation of Mn2+ into
the ceria structure was inferred. The TPR patterns were observed
to be a function of manganese content, as described before.54,56
However, at variance with the previous study, the authors here

b1469_Ch-14.indd 837 4/8/2013 12:42:34 PM


b1469 Catalysis by Ceria and Related Materials

838 A.M. Venezia et al.

concluded that there was no evidence for Mn4+ species. The agree-
ment between the analytical and the XPS-derived Mn/(Mn+Ce)
atomic ratios indicated a homogeneous distribution of manganese
up to 25 at%, thereafter segregation of the crystalline Mn3O4 phase,
as indicated by XRD, occurred. The mixed-oxide samples, tested in
the catalytic oxidation of VOCs, exhibited higher ethanol and tolu-
ene conversion (although with toluene the differences were smaller)
compared to conversion with individual oxides. Additionally, MnOx
was more active than pure CeO2. The Mn0.50Ce0.50 sample was the
most active with complete conversion to CO2 achieved at 200 °C for
ethanol and 260 °C for toluene. For ethyl acetate conversion, a simi-
lar catalyst ranking was obtained, with different reaction products
such as CO2, acetaldehyde, acetic acid, and ethanol. For ethyl ace-
tate a separate experiment with acetic acid concluded that the oxi-
dation of ethyl acetate is controlled by the reactivity of the
intermediate acetic acid, which blocks catalytic sites. Quite interest-
ingly, and in contrast to the case discussed above,54 in terms of the
specific reaction rate per unit surface area, the mixed oxides pre-
sented lower activities than single oxides.56 Moreover, the specific
activity of CeO2 for the oxidation of toluene and ethyl acetate was
higher than the activity of MnOx whereas the reverse was obtained
with ethanol oxidation. It is interesting that the enhancement of the
reducibility of the mixed oxides, as inferred from H2-TPR, did not
correspond to an increase of the specific activity in VOC oxidation.
The authors concluded that reduction in the presence of H2 as
probed by TPR was likely different when the reducing agent was a
VOC molecule. In contrast to H2, VOC molecules were adsorbed on
the catalyst surfaces under reaction conditions. This lack of correla-
tion between reducibility and VOC oxidation deserves further atten-
tion also with respect to different VOCs.
The binary system CuO–CeO2, as seen in the previous section,
is an interesting system for oxidation reactions. Different prepara-
tion procedures have been developed, enormously affecting cata-
lytic activity.28,59–62 Delimaris and Ioannides55 recently investigated
several CuO–CeO2 catalysts with different Cu/(Ce+Cu) atomic
ratios prepared by the combustion method and tested in the

b1469_Ch-14.indd 838 4/8/2013 12:42:34 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Catalysts for Air Pollution Abatement 839

oxidation of ethanol, ethyl acetate, and toluene. The combustion


method involved the use of urea as a fuel and nitrates salts as pre-
cursors, in analogy with the method used by the same authors for
the MnOx–CeO2 system.54 With increasing ceria loading the specific
surface area increased from 2.5 m2.g−1 for pure CuO to a maximum
of 53 m2.g−1 for a sample with 85 at% Ce, and back to 5 m2.g−1 when
approaching pure CeO2. According to H2-TPR, the reduction of
pure copper was characterized by two peaks, one at 268 °C attrib-
uted to the reduction of Cu2O species and the second at 323 °C
attributed to the reduction of CuO particles. As previously observed
for the mixed manganese–ceria system, the reduction peaks of the
mixed oxides were shifted to low temperatures with respect to pure
CuO. On increasing the ceria content, a broad band formed by two
overlapping peaks extending from 150 °C to 350 °C, appeared.
Quantitative analyses of the TPR patterns indicated that reduction
of ceria was occurring during the reduction of copper and that syn-
ergy between the two oxides enhanced the reducibility of both. The
XRD patterns of the binary samples were typical of the fluorite
structure of pure CeO2 with no CuO phase in the low CuO samples.
For the CuO–CeO2 catalysts with 50 and 75 at% of Cu, Cu0.5Ce0.5,
and Cu0.75Ce0.25, respectively, reflections of CuO were present in
addition to those of ceria. No variation of the ceria lattice parame-
ter was detected in any of the samples. Concerning catalytic perfor-
mance, the mixed oxides behaved much better than the pure
oxides with CuO being less active than CeO2. The positive effect of
the addition of copper to ceria increased up to a Cu content of 25
at%. A small amount of copper (5 at%) shifted the temperature for
the total conversion of ethanol from 280 °C with pure CeO2 to
230 °C. A similar trend was observed for the oxidation of ethyl ace-
tate and toluene. In contrast to the MnOx–CeO2 system yielding
only CO2 as an oxidation product of ethanol, with the CuO–CeO2
catalysts this reaction proceeded through the formation of interme-
diates: acetaldehyde for the pure oxides and ethyl acetate on the
mixed sample. Figure 14.10 shows conversion rates for the interme-
diate products of ethanol oxidation over the pure oxides and mixed
Cu0.15Ce0.85. The production of ethyl acetate indicated that the

b1469_Ch-14.indd 839 4/8/2013 12:42:34 PM


b1469 Catalysis by Ceria and Related Materials

840 A.M. Venezia et al.

Figure 14.10 Conversion rates for intermediate products of ethanol oxidation


over the catalysts CeO2, Cu0.15Ce0.85, and CuO. From Delimaris and Ioannides55 with
permission.

acetate species, formed on the catalyst surface, combined with


adsorbed ethanol in an esterification reaction in which the role of
ceria appeared crucial. Intermediate products during ethyl acetate
oxidation were acetaldehyde over pure CuO and ethanol over the
mixed oxides and both over CeO2. Whereas the effect of CO2 was
negligible for all the samples, the addition of water caused a moder-
ate decrease of the activity during ethanol and acetyl acetate oxida-
tion and a more pronounced inhibition of activity during toluene
oxidation. The increase of the surface area of the mixed oxides
compared to the pure oxides was taken into account by calculating
the specific reaction rate per unit surface area. The largest value of
the specific rate was found for pure CeO2. Thereafter there was a
decrease of specific activity with increasing surface area, which cor-
responded to an increase of the oxide crystallite sizes. This result is
indicative of a structure-sensitive reaction as obtained by the same
authors with the MnOx–CeO2 system.54 Moreover the same research-
ers reported that the specific activity of CuO–CeO2 catalysts for CO
oxidation was up to four times the value obtained with pure CuO
and this was attributed to the enhanced reducibility of CeO2.63
The apparent contradiction was suggested to be due to different

b1469_Ch-14.indd 840 4/8/2013 12:42:34 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Catalysts for Air Pollution Abatement 841

mechanisms for CO and VOC oxidation. Indeed the mechanism for


CO oxidation implied that the adsorbed CO was oxidized to CO2
through lattice oxygen. In VOC oxidation, in which catalyst reduc-
ibility did not seem to be important, the oxidation mechanism
requires several steps. As an example, for ethanol, oxidation
involves at first the adsorption of a molecule at the surface of the
catalyst and then the breaking of several C–C and C–H bonds and
one C–O bond.
The CuO–CeO2 composite oxide was also studied when sup-
ported by impregnation on inert γ-Al2O3 in the oxidation of pro-
pane and toluene.64–66 The turnover frequency in the oxidation of
propane was calculated by considering the concentration of the
active sites, which was assumed to be equal to the oxygen storage
capacity obtained from reduction/re-oxidation pulse experiments.64
A binary-oxide catalyst with molar ratio Cu/Ce ∼ 4, exhibited a
higher turnover frequency compared to the supported single
oxides, CuO and CeO2. The authors used the Mars–Van Krevelen
(MVK) mechanism to describe the kinetic data. This mechanism,
for total oxidation reactions over different metal-oxide catalysts,46,47
has a reduction step in which propane reacts with a single oxidized
surface site and a re-oxidation step with gas-phase oxygen. Applying
this mechanism to describe the kinetic data, they calculated the frac-
tion of oxidized sites and concluded that the most critical step for
propane oxidation over CuO was the reduction step whereas for the
oxidation of propane over CeO2 it was the oxidation step. Therefore
the superior catalytic activity of the binary oxide CuCeO, was attrib-
uted to the synergy between the features of both oxides and also to
the higher sticking probability for propane over the binary oxide
than over the single oxides.
The oxidation of toluene was investigated, by the same research
group, over a similar CuO–CeO2 catalyst supported on γ-Al2O3 in a
temporal analysis of products, in the temperature range 400–650 °C
in the presence and in the absence of O2 and at various degrees of
catalyst reduction.66 Once again it was shown that the total oxidation
of toluene over this type of catalyst is well described by the classical
MVK mechanism. Using isotopic labeled oxygen 18O2 it was shown

b1469_Ch-14.indd 841 4/8/2013 12:42:34 PM


b1469 Catalysis by Ceria and Related Materials

842 A.M. Venezia et al.

that adsorbed oxygen participates in the reaction. The dioxygen


ensured the re-oxidation of the reduced catalyst and also oxidized,
together with the lattice oxygen, the carbon-containing surface
intermediates between toluene and CO2. The importance of lattice
oxygen mobility and therefore of the ceria oxygen storage capacity
was clearly stated.
As recently reported by our group, a series of co-precipitated
Co3O4–CeO2 mixed oxides was studied in toluene oxidation.67 The
most efficient catalyst, with 30 wt% Co3O4, showed full toluene con-
version at 275 °C, comparing favorably with 1 wt% Pt/Al2O3 achiev-
ing 100% toluene conversion at 225 °C. According to TPR analyses,
surface and bulk oxygen species determined the activity of the com-
posite system, whereas only surface oxygen mobility governed the
activity of the single ceria oxide.

14.3.2 Ceria-supported metal catalysts


Supported noble-metal catalysts are still regarded as the most active
systems for oxidation reactions when compared to transition-metal
oxides, despite being more expensive and more sensitive to poison-
ing. A definite target of research in this field is the development of
active catalysts using a combination of transition-metal oxides with
the smallest possible amount of a precious metal.
Various CeO2-supported precious-metal catalysts, prepared by
impregnation followed by different reduction treatments, were
investigated for the catalytic combustion of ethyl acetate.68 Of the
calcined catalysts, Ru/CeO2 achieved the highest activity. Remarkably,
and in spite of the lower surface area and larger metal crystallites,
the activity of the Ru/CeO2 catalyst was also superior compared to
the activity of a Pt/γ-Al2O3 catalyst. For Pt/CeO2, Pd/CeO2, and Rh/
CeO2, catalytic activity was enhanced by a reduction treatment at
400 °C, though it was still lower compared to Ru/CeO2. The invari-
ance of Ru/CeO2 activity after reduction treatment was attributed to
the strong interaction between ruthenium species and the lattice
oxygen of CeO2 present in both pre-oxidized and pre-reduced
samples.69 H2-TPR analyses confirmed that as for the CeO2

b1469_Ch-14.indd 842 4/8/2013 12:42:34 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Catalysts for Air Pollution Abatement 843

supported precious metals, the reduction of the ruthenium species


started at the lowest temperature. The easier reducibility was likely
responsible for the higher activity in the complete oxidation of ethyl
acetate. For noble-metal catalysts, it should be recalled that the
Mars–van Krevelen mechanism, successfully used to interpret hydro-
carbon oxidation over oxide catalysts, was also applied to noble-
metal catalysts supported on an irreducible oxide like alumina. The
mechanism assumed a surface redox cycle of the catalyst metal,
alternately by oxygen and a hydrocarbon.70,71 Accordingly, the rate-
determining step during the oxidation of a hydrocarbon over Pt
catalysts was the adsorption of oxygen on metal atoms followed by
oxidation of the hydrocarbon. Although in the particular case of the
alumina-supported catalysts the role of the oxide was not important,
the combination of a noble metal and a reducible oxide like CeO2,
characterized by high oxygen surface mobility,72 was definitely ben-
eficial for hydrocarbon oxidation activity.69
When dealing with supported gold catalysts the choice of sup-
port, in contrast, is always crucial. Unlike palladium and platinum,
which are good oxidation catalysts, gold has historically been con-
sidered catalytically inactive in particular with respect to alkane
oxidation. However, as described in the previous section of this
chapter, the improvement of ceria-supported gold catalysts and the
attainment of very high activities for the low-temperature oxidation
of CO has driven interest in exploiting gold catalysts for the environ-
mental abatement of VOCs.73–76 As with the oxidation of CO, the
VOC oxidation activity of Au/CeO2 catalysts depended very much
on the preparation method, which enormously affected the size of
the gold particles and their oxidation states. In earlier work, a com-
parison between Au/CeO2 catalysts with ∼ 5 wt% Au, prepared by
deposition-precipitation and by co-precipitation, showed the superi-
ority of the former catalyst in the oxidation of a mixture of 2-pro-
panol, methanol, and toluene.73 The XRD patterns were quite
similar and the Scherrer-derived particle sizes were only slightly
different, with a smaller size for the DP sample compared to the CP
samples. XPS analyses showed metallic gold in both cases as
expected after calcination of both samples at 450 °C. However, a

b1469_Ch-14.indd 843 4/8/2013 12:42:34 PM


b1469 Catalysis by Ceria and Related Materials

844 A.M. Venezia et al.

much larger amount of gold was found at the catalyst surface of the
DP sample. The TPR patterns displayed a noticeable temperature
shift of the ceria surface peak confirming the promotion of ceria
reduction by gold. As a consequence of the better dispersion of sur-
face gold, the observed negative shifts were 260 °C and 360 °C with
the CP and the DP gold catalysts, respectively. On the basis of the
characterization results it was suggested that the main reason for the
increase of activity was the increased reducibility of the CeO2 sur-
face of the DP sample, thus enhancing the mobility of the surface
lattice oxygen involved in the VOC oxidation through the Mars–van
Krevelen mechanism.73
The same research team recently investigated a series of ceria-
supported catalysts using group IB metals (Ag, Au, and Cu) prepared
by DP and CP as above, but calcining at 300 °C rather than 450 °C.74
The combined effect of the metal and preparation method was
evaluated in the oxidation of methanol, toluene, and acetone. In all
cases the oxidation activity of ceria was highly promoted by the addi-
tion of the metal. On all catalysts VOC reactivity followed the order:
methanol > acetone > toluene. The preparation method strongly
affected the oxidation activity of the different molecules and the
effect depended on the particular metal. As shown in Fig. 14.11 for
methanol conversion, with respect to the DP samples, gold was always
the most active followed by silver and then copper. With respect to
the CP samples, silver was the most active followed by gold and cop-
per. The difference in catalyst surface area was quite small. According
to XRD, smaller CeO2 crystallites were obtained in the CP samples
compared to DP. This was attributed to the presence of the IB metal
during co-precipitation limiting ceria crystallite growth.74 According
to the H2-TPR patterns, all the metals promoted the surface reduci-
bility of the ceria, therefore increasing oxygen mobility and weaken-
ing the Ce–O bond located near the active metal atoms. In general,
the larger the metal dispersion, the larger the effect on ceria reduc-
ibility. The study established very well the dependence of the catalytic
effect on the preparation procedure of the particular metal. In par-
ticular, for co-precipitation, the solubility of the metal hydroxide with
respect to the solubility of Ce(OH)3 was very important. Due to the

b1469_Ch-14.indd 844 4/8/2013 12:42:34 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Catalysts for Air Pollution Abatement 845

Figure 14.11 Methanol conversion vs reaction temperature over catalysts pre-


pared by DP and CP. From Scirè et al.74 with permission.

b1469_Ch-14.indd 845 4/8/2013 12:42:34 PM


b1469 Catalysis by Ceria and Related Materials

846 A.M. Venezia et al.

very low solubility of Au(OH)3 (Ksp = 5×10−46) compared to Ce(OH)3


(Ksp= 1.6×10−20), gold hydroxide will precipitate before cerium
hydroxide, leading to a less intimate mixing of the components. The
opposite occurs with silver and copper, which have solubility prod-
ucts (Ksp (AgOH) = 1.2 x10−8, Ksp(Cu(OH)2) = 2x10−16) larger than ceria
hydroxide, leading to better contact or better particle dispersion. In
the presence of the support for the DP samples, the low solubility of
gold led to a large number of nucleation centers and therefore to a
better particle dispersion over the ceria. However, for the silver and
copper samples prepared by DP, the larger solubility products lead to
a lower number of nucleation sites and therefore to larger particles
in accord with the dependence of the cluster size on the ratio of
the nucleation rate to the growing rate of the metal particles.77 The
study pointed out that of the IB metal/ceria catalysts, Au/Ceria and
Ag/Ceria are the most active for VOC combustion. Considering that
Au/CeO2 (DP) had comparable activity with Ag/CeO2 (CP), the lat-
ter catalyst appears promising for practical applications.
The effect of the support on gold catalysts was recently investi-
gated in the total oxidation of 2-propanol and in the total oxidation
of propene and toluene.75,76 In both studies the superiority of gold
on ceria compared to gold over alumina or titania was shown. In the
oxidation of 2-propanol the authors observed that besides a support
effect, there was also a gold loading effect.75 Comparing gold load-
ing of 0.3 wt%, 1.6 wt%, and 2.1 wt% they found the maximum activ-
ity for 1.6 wt% gold. They explained this behavior on the basis of
reducibility, on the number of Au+1 species and on particle size,
which was below 50 Å in the lower gold content sample and around
170 Å for the 2.1 wt% sample.75 This is in accord with results for CO
oxidation, where the presence of Au+1 species seemed to be respon-
sible for the highest CO conversion at the lowest temperature.41,42
The different amount of intermediate products, propene and ace-
tone, obtained with the different gold catalysts, was attributed to
differences in the catalyst acid sites as determined by NH3-TPD pro-
files, shown in Fig. 14.12. Higher selectivity to propene was obtained
on Au/TiO2 exhibiting stronger acidic sites on its surface and more
acetone was formed with Au/CeO2 because of its basic property.

b1469_Ch-14.indd 846 4/8/2013 12:42:34 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Catalysts for Air Pollution Abatement 847

Figure 14.12. NH3-TPD profiles of catalysts: c 1.4 Au/TiO2;  1.6 Au/Al2O3; × 1.7
Au/Fe2O3; „ 1.6 Au/CeO2. From Liu and Yang75 with permission.

The low-temperature oxidation of propene and toluene on gold


catalysts (with ∼1 wt% Au) supported over CeO2, TiO2, Al2O3, and
over a ceria-promoted alumina catalyst Au/7.5Ce/Al2O3 was recently
reported by our group.76 As shown in Fig. 14.13, the trend for the
oxidation of both molecules was: Au/CeO2>Au/7.5Ce/Al2O3>Au/
TiO2>Au/Al2O3, with propene converting at a lower temperature
compared to toluene and with CO2 and H2O as the only products.
The XRD patterns of all samples showed only the crystalline features
of the CeO2 cubic structure with no gold-related peaks. The sizes of
the ceria crystallites in the alumina-supported sample were much
smaller compared to the ceria-bulk-supported catalyst. The titania-
and alumina-supported catalysts contained anatase TiO2 and the
γ-Al2O3 phase. An investigation of the chemical state by XPS showed
only metallic gold for the alumina-supported catalyst, Au+1 and Au3+
for the ceria-supported catalyst, and a mixture of metallic and
oxidized gold for the ceria-promoted alumina catalyst and in the
titania-supported catalyst. According to TEM micrographs, well
dispersed Au nanoparticles were found over CeO2, TiO2 (30–50 Å),
and also over Ce7.5/Al2O3 (20–30 Å), whereas larger particles were

b1469_Ch-14.indd 847 4/8/2013 12:42:35 PM


b1469 Catalysis by Ceria and Related Materials

848 A.M. Venezia et al.

Figure 14.13 Conversion of (a) propene and (b) toluene vs temperature over gold
supported catalysts. From Ousmane et al.76 with permission.

found over Au/Al2O3 (50–100 Å). Using a GHSV of 35000 h−1 the
Au/CeO2 catalyst was able to achieve 50% propene conversion at
152 °C and total propene conversion at 230 °C, while it achieved
50% toluene conversion at 208 °C and full toluene conversion at
290 °C.76 The efficiency of the Au/Ce7.5/Al2O3 sample was quite
interesting because only a small amount of surface-deposited ceria
was able to promote the activity of the gold over the alumina sup-
port. The supports on their own, tested in the catalytic oxidation of

b1469_Ch-14.indd 848 4/8/2013 12:42:35 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Catalysts for Air Pollution Abatement 849

propene, exhibited fair activity, especially CeO2 and ceria-impreg-


nated alumina. According to H2-TPR analyses and in agreement
with other literature data, due to the presence of gold, that part of
the ceria peak due to surface ceria in intimate contact with gold, was
shifted to a lower temperature. The surface-ceria reduction peak for
ceria not in contact with gold remained around 440 °C.76 For the
alumina-supported ceria, there was also a temperature shift of the
surface peak giving approximately the same hydrogen consumption
per gram of CeO2 as for the equivalent peak in the pure-CeO2-sup-
ported gold sample. The catalytic results shown in Fig. 14.13 con-
firm that the mechanism for the total oxidation of organic molecules
implied the activation of oxygen through a surface redox mecha-
nism, which can be promoted by gold in intimate contact with the
reducible support.
A new preparation procedure starting from the bimetallic
carbonyl cluster salt [NEt4][AuFe4(CO)16] was used to prepare
CeO2-supported Au/FeOx catalysts used for the oxidation of metha-
nol and toluene.78 Samples with different gold and iron loading
were prepared and compared with samples prepared in a similar
way but without gold. In the catalytic oxidation of methanol iron did
not affect the activity of the ceria-supported gold catalyst, in contrast
to the effect of gold, which noticeably increased the activity of bare
ceria and the FeOx/CeO2 catalysts. A correlation between the meth-
anol oxidation activity of Au/FeOx/CeO2 and catalyst reducibility as
determined by H2-TPR was observed. Quite interestingly, completely
different behavior was found for toluene oxidation. In this case bare
ceria was more active compared to FeOx/CeO2 or Au/FeOx/CeO2.
This was explained as being due to the different interaction between
the organic substrate and the catalytic surface. Based on TPD exper-
iments, whereas alcohols are strongly adsorbed on the basic sites of
metal oxides such as iron or cerium with the formation of oxygenate
species, the interaction of toluene with the oxide surface sites was
much weaker.78,79 Therefore the activation of an organic molecule,
which is important for its oxidation, is more difficult over FeOx/CeO2
catalysts due to the partial coverage of the ceria sites, considered
responsible for toluene adsorption48 by the less adsorptive FeOx.79
The presence of gold somehow increased the activity due to

b1469_Ch-14.indd 849 4/8/2013 12:42:35 PM


b1469 Catalysis by Ceria and Related Materials

850 A.M. Venezia et al.

increased oxygen mobility but did not compensate for the loss of
ceria adsorption sites. The authors concluded that the positive effect
of gold was determinant under those conditions where the activa-
tion of the surface oxygen is the key step for deep oxidation reac-
tions as in methanol oxidation.
As shown from the literature results mentioned above, despite
the significant progress in the development of catalysts for the total
oxidation of organic molecule, it is actually very difficult to design a
system that works efficiently as a general VOC combustion catalyst.
The variety of molecules in a real feed requires specific catalysts for
different groups of compounds, such as aliphatic, aromatic, and
chlorinated.

14.4 Methane Combustion


Due to the strong greenhouse effect of this gas, increasing concern
about environmental pollution demands highly efficient catalysts
for the complete abatement of methane emission. The total oxida-
tion of methane occurs in catalytic combustors for power genera-
tion, involving high combustion temperatures, and in catalytic
converters for lean-burn natural-gas vehicles (NGVs) working typi-
cally at temperatures of 320–420 °C.80,81 In the latter, without any
post-treatment device, the amount of unburned methane released to
the atmosphere exceeds the limit of 1.6 g.kWh−1 in Euro III and
1.7 g.kW−1 in US standards.80 It has been established that a catalyst
able to treat exhaust gases of natural gas vehicles must have the fol-
lowing characteristics: high efficiency for a very diluted feedstream
(less than 1000 ppm of methane) with high oxygen concentration,
low light-off temperature at high space velocities, high thermal
stability, and resistance to deactivation by water and sulfur oxides
(SOx). In contrast to the oxidation of VOCs, the development of
efficient catalysts for the total oxidation of methane appears to be a
simpler task, dealing essentially with only one substrate. However,
the challenge resides in the intrinsic difficulty of activating the CH4
molecule. Both classes of catalysts, metal oxides82,83 and noble
metals,84–86 are currently being explored. The former class has the

b1469_Ch-14.indd 850 4/8/2013 12:42:35 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Catalysts for Air Pollution Abatement 851

advantage of being less expensive while the latter is more thermally


stable, is generally more efficient especially at low temperatures, and
has a greater resistance to sulfur poisoning although it has higher
volatility and poor availability.87,88

14.4.1 Ceria-based oxide catalysts


The complete oxidation of methane into water and carbon dioxide
over a ceria surface has been investigated by a computational
method aiming to understand the suitability of CeO2 as a possible
anode material for solid oxide fuel cells (SOFCs).89 Calculations
were performed using density functional theory (DFT) for the (111)
surface of CeO2, the surface of ceria which is the most stable and the
hardest to reduce. Of the different possible pathways for the abstrac-
tion of the first hydrogen atom the most favored is the so-called
“rebound mechanism” in which methane is adsorbed through the
interaction between hydrogen and ceria surface oxygen, leaving a
methyl radical, which undergoes successive dissociation. Activation
energies were calculated for all dissociation steps and, with a barrier
of 33 kcal.mol−1, the abstraction of the first hydrogen atom was
shown to be the rate-limiting step. Quite importantly from the SOFC
perspective, the complete dissociation of methane over a ceria sur-
face generated CO2 without adsorption of carbon and the conse-
quent formation of carbon coke. Despite these encouraging results,
from a practical point of view, ceria on its own is much less active in
the combustion of methane than other transition-metal oxides like
CuO, Co3O4, etc. Its easy thermal sintering and its basic character
are important drawbacks, which hinder catalytic efficiency.
In our group we explored the combination of CeO2 with Co3O4
to obtain active methane combustion catalysts.9,82 Based on their rela-
tive methane oxidation activities, CeO2 was considered as the sup-
port whereas Co3O4 was the active phase. Samples of Co3O4/CeO2
with different amounts of Co3O4 were prepared by co-precipitation
from nitrate precursors in the presence of citrate at pH = 9 and by
impregnation of commercial CeO2 with an aqueous solution of
cobalt nitrate. In both cases the samples underwent final calcination

b1469_Ch-14.indd 851 4/8/2013 12:42:35 PM


b1469 Catalysis by Ceria and Related Materials

852 A.M. Venezia et al.

at 650 °C. The Co3O4/CeO2 catalysts prepared by co-precipitation


and impregnation exhibited superior activity compared to the single
oxides. The best catalyst was 30 wt% Co3O4 with a methane conver-
sion of 50% at around 450 °C, while over CeO2 half methane conver-
sion was at around 580 °C. After thermal aging at 750 °C for 7 h, the
co-precipitated sample maintained its rate of methane conversion
whereas the impregnated sample was strongly deactivated. XRD
analyses of the fresh and spent samples indicated, to a large extent,
sintering of the CeO2 and Co3O4 particles of the impregnated sample
on aging. TPR analyses of the aged samples indicated a better reduc-
ibility of the co-precipitated sample with respect to the impregnated
one. Moreover, as a way to increase the thermal stability of the car-
rier, a mixed CeO2–ZrO2 oxide was also used as a support. Co3O4/
CeO2 and Co3O4/CeO2–ZrO2 catalysts, both prepared by co-precipi-
tation, were compared for methane combustion activity using differ-
ent reaction conditions.83 The pure oxides CeO2 (CeO2prec) and
CeO2–ZrO2 (CeZrcopr) with nominal composition Ce0.6Zr0.4O2 were
also prepared by precipitation with sodium carbonate. Figure 14.14
shows methane conversion curves as a function of temperature for
fresh and aged samples containing 30 wt% of Co3O4, for the reaction
performed in lean conditions corresponding to λ = 8 and rich condi-
tions corresponding to λ = 1. The parameter λ is defined as the con-
centration of O2 in the feed divided by the stoichiometric O2 needed
to oxidize CH4. As shown in the figure, the fresh samples behaved in
a similar way in excess oxygen (λ = 8) except that CeO2 and CeZr are
much less active compared to the samples containing cobalt. The
extent of the catalytic decay of the aged catalysts depended on the
type of sample, with pure Co3O4 exhibiting the largest degree of
deactivation. In stoichiometric conditions (λ=1) catalytic conversion
decreased in all samples as indicated by the higher temperature shift
of the curves, and the largest deactivation of an aged sample was
observed with pure Co3O4. The thermal instability of the Co3O4
active phase was shown by the decrease of activity at around 800 °C
because of the decomposition of cobalt oxide. The stabilizing effect
of the presence of ceria and zirconia was reflected in the preserva-
tion of activity at this high temperature. Structural characterization

b1469_Ch-14.indd 852 4/8/2013 12:42:35 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Catalysts for Air Pollution Abatement 853

Figure 14.14 CH4 conversion as a function of temperature over fresh (filled sym-
bols) and aged (open symbols) samples: {z Co3O4prec; dV Co30Cecopr; cU
Co30CeZrcopr;  CeZrcopr; Ë CeO2prec; at (a) λ = 8, (b) λ = 1, with WHSV = 60000
mL.g−1.h−1. From Liotta et al.83 with permission.

b1469_Ch-14.indd 853 4/8/2013 12:42:35 PM


b1469 Catalysis by Ceria and Related Materials

854 A.M. Venezia et al.

by XRD, TPR, and XPS confirmed the important role of the ceria
and ceria-zirconia support in maintaining the reducibility of the sam-
ple and also in avoiding sintering of the Co3O4 particles.83

14.4.2 Noble-metal catalysts


Noble-metal catalysts are more efficient and can achieve total
methane conversion at lower temperatures (below 300 °C) and are
used in after-treatment devices for natural gas vehicles. Palladium,
platinum, or rhodium and alloys of these metals are generally
used. Palladium is one of the most active precious metals in the
total oxidation of methane, although it is sensitive to traces of
water and SOx. For this type of catalyst a key role is played by the
dynamic equilibrium between the reduced and oxidized states of
palladium, Pd and PdO, and by the support. The most common
supports are alumina, silica, and zirconia. In order to enhance
catalytic efficiency, rare-earth oxides (REO) are typically used as
structural and electronic promoters.90 CeO2 is the most common
promoter since, as mentioned above, it enhances the dispersion of
the noble metal, promotes noble-metal oxidation and reduction
under reaction conditions, and increases the thermal stability of
the support.90,91 The potential of CeO2 as a catalyst component for
methane combustion has recently been investigated.91–93 The acti-
vation barrier for methane over surfaces of CeO2(111), PdO(100),
and palladium-incorporated ceria, PdxCe1-xO2(111), were com-
pared through a computational study using density functional
theory.94 As found for pure CeO2,88 the rate-determining step in
the total oxidation of methane is C–H activation for which the
energetic barrier over PdxCe1-xO2(111) was the lowest compared to
the other surfaces, Pd(111), stepped Pd(211), PdO(100), Pd/
CeO2(111), and CeO2(111). The study indicated that Pd4+ ions
incorporated in ceria provided higher catalytic activity for meth-
ane oxidation than large palladium particle, Pd(111), or large pal-
ladium oxide particle, PdO(100). The calculated rate of methane
oxidation was three orders of magnitude larger than over stepped
Pd(111) and much larger than over Pd(111), CeO2(111), and

b1469_Ch-14.indd 854 4/8/2013 12:42:35 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Catalysts for Air Pollution Abatement 855

PdO(100). The theoretical study was in agreement with experi-


mental studies establishing the importance of the Pd–O–Ce link-
ages and more generally oxidized Pd species in dictating the
catalytic activity of Pd/ceria. The model illustrated how mixed
noble-metal–ceria oxides could provide enhanced activity for
hydrocarbon oxidation by tuning the reducibility of the surface.89
A study carried out in our group has shown how the addition of
0.7 wt% Pd to mixed Co3O4–CeO2 and to CeO2 improved methane
oxidation activity by lowering T50 by 100 °C.95 Palladium catalysts for
the oxidation of methane suffer the drawback of being easily poi-
soned by sulfur compounds.85,86 Indeed for Pd/Co3O4 a mild and a
drastic deactivation by sulfur poisoning was observed under both
stoichiometric and lean conditions.96 Adding CeO2 to the support
allowed the maintenance of stable activity in the presence of SO2
both in lean and in stoichiometric conditions.95 Moreover, as shown
by XPS analyses, the presence of ceria prevented the complete
reduction of palladium by SO2, which, in contrast, occurred for the
Pd/Co3O4 catalyst.
In a comparative study of the effect of different rare-earth oxides
such as Tb4O7, Pr6O11, and CeO2, on the methane combustion activ-
ity of a γ-Al2O3-supported palladium catalyst, CeO2 positively affected
the stability of the active species PdO and consequently combustion
activity.90 In this study, the supports with 15 wt% REO were prepared
by impregnating γ-alumina with a rare-earth nitrate solution followed
by drying and calcination at 800 °C. Thereafter palladium (10 wt%)
was supported by impregnation with a Pd(NO3)2 solution, again fol-
lowed by drying and calcination at 800 °C. After high-temperature
calcination, Tb4O7, and Pr6O11 conferred thermal stabilization
against the loss of alumina surface area probably due to the forma-
tion of aluminate on the alumina surface, which would lock highly
reactive coordinated unsaturated Al ions.90 In contrast, CeO2 did not
affect the thermal stability of the support, probably due to the fact
that it did not form a binary oxide. The stabilization of Pr and Tb as
aluminates influenced their redox properties making them ineffec-
tive in promoting PdO reformation during combustion. The effect
of an REO on the active species, considered to be PdO, and

b1469_Ch-14.indd 855 4/8/2013 12:42:35 PM


b1469 Catalysis by Ceria and Related Materials

856 A.M. Venezia et al.

therefore on the activity, was evaluated by performing temperature-


programmed oxidation (TPO) cycles and temperature-programmed
methane combustion experiments. According to the hysteresis of
the curves obtained with all three REO-doped catalysts, CeO2 was
the only dopant that enlarged the PdO stability window. In accord
with previous studies on alumina-supported catalysts, PdO started to
decompose to metallic Pd even in the presence of oxygen at tem-
peratures above 670 °C, then during the cooling cycle Pd re-oxidized
at temperatures around 470 °C. In the presence of CeO2, the decom-
position and the re-oxidation temperatures were closer to each
other. The hysteresis of Pd–PdO was reduced to ca. 60 °C instead of
150 °C for the unpromoted sample, with a consequent increase of
the temperature range for PdO stability. In a recent study by the
same group, on the basis of XPS analyses it was established that PdO
decomposition and Pd re-oxidation occurred via the formation of
intermediates consisting of surface or interfacial PdOx.97
Simplício et al. studied the effect of ceria modulated by the effect
of palladium precursors on methane oxidation activity of supported
palladium catalysts.98 Different PdO/Al2O3 and PdO/CeO2/Al2O3
catalysts with Pd loading near 3 wt% were prepared by impregnation
of the support with a solution of Pd(NO3)2, Pd(C5H7OO)2, and PdCl2
as precursors. The ceria-containing support was prepared by impreg-
nation of alumina with an aqueous solution of Ce(NO3)3 to yield a
final CeO2 loading of 10 wt%. All the samples underwent a calcina-
tion treatment at 1000 °C in air flow. Metallic Pd and PdO along with
γ-Al2O3 and the fluorite CeO2 phase were detected by XRD. The
thermal stability studied by TPO agreed with the previous studies
showing the stabilizing effect of ceria on the PdO phase. During the
heating ramp, PdO decomposition peaks were observed between
780 and 840 °C, little influenced by the palladium precursors and by
ceria doping. In contrast, during the cooling cycle the oxygen con-
sumption peak corresponding to the re-oxidation of the samples
was strongly affected by the palladium precursors and the presence
of ceria. The effect of ceria in enhancing PdO thermal stability
was indicated by the high-temperature shift of the palladium re-
oxidation peak. The extent of the effect was modulated by the type

b1469_Ch-14.indd 856 4/8/2013 12:42:35 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Catalysts for Air Pollution Abatement 857

of the palladium precursors, which ultimately determined the PdO–


CeOx interaction. The presence of ceria and a favorable PdO–CeOx
interaction had a strong impact on the methane oxidation activity of
the catalysts. Comparing the curves on the left side of Fig. 14.15 with
those on the right side, it is evident that ceria limited the loss of activ-
ity during the cooling cycle due to PdO decomposition. Moreover
comparing the result on the ceria-promoted catalysts, on the left
side, obtained with two different palladium precursors, the loss of
activity was much less in the case of acetylacetonate compared to the
nitrate precursor. This result was in agreement with the stronger
Pd–Ce interaction in the former sample as was shown by Fourier
transform infrared (FTIR) spectroscopy by Monteiro et al.99

Figure 14.15 Methane conversion obtained with (A) Pd/Al2O3 and (B) Pd/CeO2/
Al2O3 from Pd(C5H7OO)2 (top) and Pd(NO3)2 (bottom). From Simplício et al.98
with permission.

b1469_Ch-14.indd 857 4/8/2013 12:42:35 PM


b1469 Catalysis by Ceria and Related Materials

858 A.M. Venezia et al.

In a more recent study the effect of ceria loading for alumina-


supported palladium catalysts was investigated.91 Supports with 2, 5,
10, 15, and 50 wt% of ceria were prepared by the sol-gel method
starting from the precursors aluminum alkoxide and cerium
nitrate. Thereafter palladium in quite small amounts (0.1–0.2 wt%)
was added by impregnation followed by calcination at 600 °C. XRD
and surface area analyses of the supports showed essentially amor-
phous patterns and quite high surface area (above 200 m2.g−1) for
the supports with a ceria content up to 15 wt%. A well-crystallized
fluorite structure of CeO2 and lower surface area were observed for
the 50 wt% ceria loaded sample. From TPR analyses it was shown
that, compared to the pure supports, the palladium enhanced the
reducibility of ceria. In the catalytic oxidation of methane under
stoichiometric conditions the best catalyst with the lowest light-off
temperature was the sample with 15 wt% of ceria. The least active
catalyst was the sample containing the highest amount of ceria. The
behavior was explained as a synergistic effect between the noble
metal and non-stoichiometric ceria with larger oxygen mobility,
favoring a partial reduction of the precious metal in the bulk of
PdO providing the Pd/PdO active ensemble sites. The detrimental
effect with a large amount of ceria was attributed to an excessive
oxidation of PdO and to a likely encapsulation of the noble metal
by CeO2.91
Compared to palladium catalysts, platinum-based systems are
more sulfur tolerant. Unlike the former, having PdO as the active
phase, platinum catalysts have metallic platinum as the active spe-
cies. As a consequence, with this type of catalyst the methane oxida-
tion rate was found to be strongly dependent on the stoichiometry
of the reactant gas mixture, being generally inhibited under an
excess of oxygen, as was the case for Pt/Al2O3.100 One way to circum-
vent the problem is to feed the gas alternating between a net-
oxidizing and a net-reducing composition so that temporarily high
oxidation activity is associated with changes in gas composition.
Becker et al. showed that the period with high activity could be pro-
longed by using oxygen-buffering ceria as a support material.101
As discussed previously, the use of ceria increased the sulfur tolerance

b1469_Ch-14.indd 858 4/8/2013 12:42:35 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Catalysts for Air Pollution Abatement 859

of supported palladium catalysts.95,96 With platinum catalysts, meth-


ane oxidation activity was even promoted by the presence of SO2 in
the reaction mixture.100 The influence of SO2 on methane oxidation
over 4 wt% Pt/CeO2 catalyst was studied under lean conditions by a
transient flow reaction and in situ FTIR spectroscopy. The variation
in methane and CO2 concentration during SO2 exposure at 400 °C
was measured.100 A decrease in methane concentration and a conse-
quent increase in CO2 concentration corresponding to SO2 expo-
sure were observed. As soon as the SO2 feed was stopped, the CH4
concentration started to increase. According to DRIFTS results,
sulfates formed on the ceria immediately after SO2 exposure and
were accompanied by the formation of Ce3+. The instantaneous
increase in methane conversion was attributed to the ability of the
reduced ceria to withdraw oxygen from platinum thereby increasing
the number of available sites for the dissociative adsorption of meth-
ane and also preventing oxidation of the platinum sites. However, as
a function of time on stream the promotional effect of SO2 dimin-
ished and was replaced by catalyst deactivation. This detrimental
effect was attributed to decreased oxygen mobility in the ceria sup-
port when completely saturated with sulfates. As shown in a previous
study by the same group, thermal regeneration of the Pt/CeO2 cata-
lyst from SOx adsorption occurred at different temperatures depend-
ing on platinum loading. The higher the amount of platinum the
lower the SOx desorption temperature.102
Gold catalysts on suitable supports have been also proposed for
methane oxidation. A comparative study of the oxidation of CO and
CH4 over Au/MOx/Al2O3 (with M = Cr, Mn, Fe, Co, Ni, or Zn) cata-
lysts prepared by deposition-precipitation established that whereas
in CO oxidation the activity correlated essentially with gold particle
size for the oxides, in the case of methane oxidation, the activity
depended on both gold particle size and the identity of the MOx.103
Quite a long time ago, in a study of the activity of co-precipitated
gold catalysts for methane oxidation it was shown that Au/CeO2 was
the least active and Au/Co3O4 the most active catalyst.104 A dual-site
mechanism and also a relation with the oxidation state of gold was
proposed to explain the synergism of gold with some of the oxides.

b1469_Ch-14.indd 859 4/8/2013 12:42:35 PM


b1469 Catalysis by Ceria and Related Materials

860 A.M. Venezia et al.

However, the inactivity of the 5 wt% Au/CeOx catalyst was not


understood since unsupported CeOx was reported as being moder-
ately active toward methane oxidation. Based on the good methane
oxidation activity exhibited by the mixed-oxide system Co3O4–CeO2
with the Co/Ce atomic ratio equal to 1,82,105 the reciprocal effect of
gold and a mixed oxide was investigated for an Au/Co3O4–CeO2
sample (Co/Ce atomic ratio = 1) prepared by co-precipitation.106
For comparison purposes, catalysts of gold on single oxides, Au/
CeO2 and Au/Co3O4, were prepared by the same method. The cata-
lytic tests, performed in lean conditions, were in accord with the
earlier literature,104 with gold on Co3O4 being much more active
than Au/CeO2. For the three samples, the effect of gold on the
active supports was an enhancement of the activity, which was more
substantial for the cobalt oxide catalyst. The activity trend Au/
Co3O4>Au/Co3O4–CeO2>Au/CeO2 matched the activity trend of
the corresponding oxides. The ceria component had an explicit
effect at temperatures above 600 °C. Indeed in these conditions the
activity of Au/Co3O4 decreased because of Co3O4 decomposition
whereas the activity of Au/Co3O4–CeO2 continued to increase
reaching full methane conversion. As discussed before, structural
promotion was due to enhanced redox properties and also to the
particle size stability enhanced by the mixed-oxide system.105
Furthermore, the methane oxidation activity of the two catalysts was
measured in the presence of 10 ppm of SO2 in order to simulate
more realistic operating conditions. Methane conversion curves for
Au/Co3O4 and Au/Co3O4–CeO2 without SO2 and with SO2 are
shown in Fig. 14.16. The positive effect played by the ceria with
respect to SO2 tolerance is obvious. Whereas the gold over Co3O4
was strongly deactivated by SO2, the mixed-oxide-supported gold
maintained reasonably good activity. The different behavior exhib-
ited by the two catalysts, in analogy with a previous study on palla-
dium supported over similar oxides,95 was ascribed to the basic
nature of ceria, which, forming a stable cerium sulfate, acted as a
chemical trap, quickly removing SO2 from the feedstream. As a con-
sequence, the presence of ceria prevented the irreversible poisoning
of the cobalt or of the gold phase.

b1469_Ch-14.indd 860 4/8/2013 12:42:36 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Catalysts for Air Pollution Abatement 861

Figure 14.16 Methane conversion vs temperature over Au/Co3O4 and Au/Co3O4–


CeO2 catalysts in the presence of 10 ppm of SO2 in the feed. From Liotta et al.106
with permission.

14.5 Reduction of NO with CO


NO reduction by CO is a primary reaction occurring in three-way
catalysts (TWC), which in a range of air-to-fuel ratios close to stoi-
chiometric mixtures, allows, at least from thermodynamic argu-
ments, the removal of three pollutants, namely NOx, CO, and HC,
leaving only H2O, CO2, and N2.107 Typically TWCs are composed of
noble metals (Pt, Rh, and Pd) dispersed on alumina. However, the
nature of the support also has an important role in affecting the
selectivity of the reactions. The reduction of NO by CO can be
written as

2CO + 2NO → N2 + 2CO2 (14.1)

It can also be shown as a two-step reaction involving the forma-


tion of the intermediate species N2O according to the following
sequence:

b1469_Ch-14.indd 861 4/8/2013 12:42:36 PM


b1469 Catalysis by Ceria and Related Materials

862 A.M. Venezia et al.

CO + 2NO → N2O + CO2 (14.2)

N2O + CO → N2 + CO2 (14.3)

Current TWC devices are able to remove exhaust NOx for


engines operating close to stoichiometric conditions, but they are
not adequate at low temperature and in lean conditions. One draw-
back is the incomplete reduction of NO into the undesirable N2O,
particularly during the cold-start regime.107,108 Stringent regulations
require more efficient noble-metal catalysts, which are able to selec-
tively convert NO to N2 at low temperatures or enhance the reaction
N2O + CO according to Eq. (14.3). In order to understand the effect
of the support in TWC and the role of CeO2 in this type of reaction,
a short description of the interaction mechanism of NO with the
main active site of the metal is given as follows.
In situ infrared spectroscopy has been widely used to investigate
intermediates and to determine the influence of the oxidation state
of rhodium and palladium in TWC.109,110 After NO exposure, posi-
tively charged, neutral, and negatively charged NO species on Rh
were detected.109 NO has an electronic structure very similar to that
of CO and the corresponding molecular orbitals (2π and 5σ)
involved in the formation of the metal–NO bond are the same. The
only difference is that NO possesses an unpaired electron in the
antibonding 2π orbital, which is the reason for the higher probabil-
ity for the dissociation of NO compared to CO. The cationic species
(Rh–NO+) is produced by the donation of an electron from the
antibonding π*2pz orbital of NO to the d orbital of the metal to
strengthen the N–O bond, while the anionic species (Rh–NO−) is
produced by the transfer of a d electron of the metal to the antibond-
ing π*2pz-orbital of NO to weaken the N–O bond. Due to these dif-
ferences in the characteristics of the bonding of the two species,
doping of the support was expected to affect, in a different manner,
the positively and negatively charged species.110 It was generally
accepted that NO breaking occurred easily over bent Rh–NO−
species and that N2O formed on oxidic rhodium. A correlation of
catalytic and infrared spectroscopy measurements, particularly the

b1469_Ch-14.indd 862 4/8/2013 12:42:36 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Catalysts for Air Pollution Abatement 863

disappearance of the Rh–NO+ band (1,869 cm−1) and the appear-


ance of the band due to N2O (at 2,243 cm−1), indicated that N2O
originated from Rh–NO+, while neutral or negatively charged NO
species, which dissociated more easily, led directly to the production
of N2. The negatively charged species dissociated to form nitride
and oxygen atoms. In the absence of an oxygen-atom scavenger, the
catalyst surface becomes progressively oxidized due to the accumu-
lation of oxygen atoms, thus favoring the formation of the less
reactive Rh–NO+ species.110

14.5.1 Ceria-promoted metal-supported catalysts


Based on the above remarks, it is conceivable that CeO2 can be used
as a suitable promoter of TWCs mainly because of its oxygen storage
capacity. The influence of the Ce additive on the catalytic perfor-
mance of a three-way bimetallic catalyst Pt–Rh/Al2O3 in the CO+NO
reaction was investigated by Granger et al.111 The ceria-doped sup-
port was prepared by impregnation of γ-Al2O3 with an aqueous solu-
tion of cerium nitrate followed by drying and calcination at 550 °C.
Metal deposition was carried out by co-impregnation followed by
calcination at 350 °C and H2 reduction at 450 °C. A beneficial effect
of ceria on the conversion of NO was observed mainly at low tem-
peratures.111 A comparison between the temperature-programmed
conversion curves for the CO+NO reactions on Pt–Rh/Al2O3 and
Pt–Rh/Al2O3–CeO2 for the first and second runs are shown in
Fig. 14.17. The conversion of NO with Pt–Rh/Al2O3–CeO2 started at
a temperature below 50 °C instead of the 150 °C for the less active
alumina-supported catalyst. Then the ceria-doped sample reached a
steady conversion at 200 °C and thereafter its activity increased again
after 280 °C. Above this temperature the two catalysts exhibited the
same behavior with superimposition of the curves suggesting a simi-
lar mechanism for the two samples and in particular that, at the
higher temperature, the mechanism occurring on Pt–Rh/Al2O3 was
dominant also on the ceria-doped sample. The catalytic test was
continued on Pt–Rh/Al2O3–CeO2 at 500 °C for 72 h and then there
was a descending temperature run, cooling the reactant mixture

b1469_Ch-14.indd 863 4/8/2013 12:42:36 PM


b1469 Catalysis by Ceria and Related Materials

864 A.M. Venezia et al.

Figure 14.17 Conversion curves for the CO+NO reaction on Pt–Rh/Al2O3 (∆) and
on Pt–Rh/Al2O3–CeO2 („ first run); (… second run after 72 h aging in the course
of the reaction at 500 °C, with PNO = PCO= 5×10−3 atm, mcat = 0.2 g and total flow rate
= 10 L.h−1). From Granger et al.111 with permission.

down to room temperature. In the descending cycle conversion in


the low-temperature range was observed again, but deactivation of
the catalyst had occurred. The low-temperature behavior with the
enhancement of activity by the presence of ceria was explained by a
bi-functional mechanism. This mode involved reaction of the oxy-
gen from ceria with adsorbed CO molecules on the metals at the
metal/ceria interface, with the consequent formation of anionic
vacancies being oxidized by the dissociation of NO adsorbed on the
metals. The low concentration of anionic vacancies prevents a
strong interaction between the ceria and the noble metals, preserv-
ing the adsorption properties of the Pt–Rh/Al2O3 catalyst. From the

b1469_Ch-14.indd 864 4/8/2013 12:42:36 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Catalysts for Air Pollution Abatement 865

temperature dependency of the rate constants, the activation energy


corresponding to the step in oxygen consumption by the adsorbed
CO was found to be quite higher than the activation energy for the
refilling step of anionic vacancies. Consequently, at high tempera-
tures the extent of ceria reduction is too high for the bi-functional
mechanism and therefore the synergy between the support and the
metal disappears. The deactivation observed after the sustained
treatment at 500 °C was explained by the extensive reduction of
ceria to Ce3+ with the consequent formation of anionic vacancies in
the vicinity of the noble metals. The consequent strong metal/
support interaction, according to spectroscopic characterization,
decreases the adsorption property of the metals and attenuates the
beneficial effect of ceria on NO conversion.
A bi-functional mechanism was previously proposed for
ZrxCe1-xO2 mixed oxides promoted with palladium.112 The substan-
tial enhancement of activity was attributed to the involvement of
active sites on Pd and ceria. Also in this case, deactivation of the cata-
lysts at a high temperature was observed and attributed to an
extended reduction of the ceria, which was not compensated by fast
re-oxidation.

14.5.2 Ceria-supported metal and oxide catalysts


Besides being used as a promoter of TWCs, ceria has also been inves-
tigated as a support of the noble metals in a TWC. One important
aspect of ceria-supported catalysts is the interaction between the
metal and the support, which can be optimized by an appropriate
choice of preparation procedure.107 The combustion method, start-
ing from (NH4)2Ce(NO3)6 and H2PtCl6 with oxalyl dihydrazide used
as fuel, was used to prepare a series of Pd/CeO2 and Pt/CeO2 cata-
lysts with different metal loading, tested, among others, in the
NO+CO reaction. From XRD, TEM, and XPS results it was ascer-
tained that with the low metal loading (1% by mole), Pt2+ and Pd2+
were incorporated into the CeO2 matrix as a solid solution unlike
similarly prepared alumina catalysts, which have separate platinum
and palladium phases. The enhancement of catalytic activity of

b1469_Ch-14.indd 865 4/8/2013 12:42:36 PM


b1469 Catalysis by Ceria and Related Materials

866 A.M. Venezia et al.

ceria-supported catalysts compared to alumina catalysts was attrib-


uted to the enhanced metal dispersion and mostly to the electronic
interaction involving the orbitals of the ionic species Ce4+(4f),
O2−(2p), Pd2+(4d), and Pt2+(5d), which facilitated redox
reactions.107
Ceria was also studied as support of base metals such as copper-
based catalysts. Cu/CeO2 and Ce1-xCuxO2 received special attention
due to the strong interaction between copper and ceria and to their
strong activity in several reactions. The synergy between copper and
ceria and the role of the different copper species in the reduction
of NO by CO was investigated by Zheng’s group in China.113 They
prepared Cu/CeO2 and Cu/SiO2 by impregnation and, in order to
detect different copper species, they removed CuO crystallites dis-
persed on the support surface by treating the catalysts in a HNO3
solution and compared these samples with untreated ones. Using
several characterization techniques including XRD, electron para-
magnetic resonance, and Raman spectroscopy, they found that after
nitric acid treatment of the Cu/CeO2 catalyst, there was only a small
amount of copper left (0.27 wt% Cu vs 5 wt% Cu in the fresh
sample). The copper was present in different forms: (i) isolated cop-
per ions in octahedral sites of ceria, (ii) copper oxide clusters, and
(iii) copper ions in the ceria lattice (Cu–O–Ce solid solution).
Moreover NO-TPD measurements showed that the active sites for
NO dissociation were the oxygen vacancies on the ceria and the
metallic copper, with the oxygen vacancies on ceria being more
active. As suggested by TPR analyses, oxygen was likely transferred
from ceria to copper particles forming metastable copper oxide,
known to be very active in CO oxidation. The trapping of the oxy-
gen by CO in copper sites maintained ceria in a reduced state, con-
sidered to be the active sites for NO dissociation, and therefore led
to high catalytic activity in the NO+CO reaction. This is shown by
the NO conversion curves of Fig. 14.18 for different samples initially
reduced in 1% CO/He at 300 °C for 1 h. Pure CeO2 showed moder-
ate NO conversion at 400 °C. Both Cu/CeO2 and Cu/CeO2–N
(treated with HNO3) exhibited much higher activity than Cu/SiO2.
The positive effect was indeed attributed to the additional

b1469_Ch-14.indd 866 4/8/2013 12:42:36 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Catalysts for Air Pollution Abatement 867

Figure 14.18 NO conversion in the NO + CO reaction using 5000 ppm NO and


5000 ppm CO in He stream at GHSV = 32000 h−1: • Cu/CeO2; d Cu/CeO2–N; c
Cu/SiO2; „ CeO2; _ Cu/SiO2–N. From Chen et al.113 with permission.

dissociation of NO on ceria vacancies, which refilled the oxygen


vacancies essential for the maintenance of catalytic activity in the
NO+CO reaction. The same group had previously reported the posi-
tive effect of adding MgO to Cu/CeO2 catalysts. Besides an increase
in surface area and CuO surface dispersion, the main reason for the
promotion was attributed to the formation of the non-equilibrium
MgxCe1-x/2O2 solid solution with more oxygen vacancies leading to
an increase in ionic mobility, therefore promoting formation of the
Cu–O–Ce solid solution.114
Ceria-supported NiO catalysts have also been investigated for
stoichiometric NO/CO and NO/CO/O2 reactions.115 Nickel in a
nominal amount of ca. 7 wt% was deposited by wetness impregna-
tion onto different CeO2 supports. The superior performance of
NiO/CeO2 was established with respect to samples of bare CeO2 and
samples of NiO supported over TiO2 and Al2O3. An important effect
of the CeO2 preparation method on catalytic activity was observed.
Of the different methods of synthesis, homogeneous precipitation
with urea produced the best catalyst. The interaction between
nickel and ceria was particularly crucial for catalytic behavior. The

b1469_Ch-14.indd 867 4/8/2013 12:42:36 PM


b1469 Catalysis by Ceria and Related Materials

868 A.M. Venezia et al.

impregnation of the ceria powder with a nickel precursor solution


allowed incorporation of Ni2+ ions into the ceria lattice. Due to the
smaller radius of Ni2+ (0.072 nm) compared to Ce4+ (0.101 nm) and
also to the different oxidation states of the two ions, charge com-
pensation and lattice strain produced by the introduction of Ni2+
into the CeO2 lattice created oxygen vacancies on which the
adsorption of NO was favored. In a subsequent paper, through the
use of FTIR and mass spectroscopy, the same group was able to
predict the mechanism for the reduction of NO in the presence of
CO.116 By performing stepwise adsorption of CO and NO in differ-
ent sequences and co-adsorption of both molecules, unlike what
was observed for noble-metal catalysts110 it was concluded that NO
adsorbed associatively over the Ni2+ of the catalyst. N2O was then
generated by the decomposition of hyponitrite N2O2 formed by the
adsorption of two molecules of NO located closely at a Ce3+–oxygen
vacancy site. The addition of CO to a surface pre-treated with NO
generated only CO2 and NO indicating that the adsorbed NO
could not interact directly with CO to produce N2. In contrast, on
a catalyst surface previously treated with CO, due to the creation of
oxygen vacancies after the formation of CO2 the addition of NO
produces N2 and NCO. The adsorbed isocyanate was thermally
unstable giving rise to CO2 and N2. Moreover, such adsorbed spe-
cies easily react with gaseous NO again producing CO2 and N2.
Similar surface species were generated by the co-adsorption of NO
and CO. According to mass spectroscopy, gaseous N2 and CO2 were
released. On the basis of these results, the reaction steps of NO
reduction by CO over NiO/CeO2 are: (i) removal of surface oxy-
gen by CO to create vacant sites; (ii) NO dissociation to produce N2
on the vacant sites; (iii) replenishment of the oxygen from NO
dissociation.
As discussed in Sections 14.1 and 14.2 of this chapter, nanostruc-
tured gold catalysts exhibited surprisingly high activity in low-tem-
perature oxidation reactions. Some earlier studies reported also on
gold activity for the reduction of NO with CO, H2 and hydrocar-
bons.117 Although gold catalysts were not able to match the high-
temperature performance of TWCs using platinum group metals,

b1469_Ch-14.indd 868 4/8/2013 12:42:36 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Catalysts for Air Pollution Abatement 869

the peculiarity of being active at a relatively low temperature is an


important incentive to investigate gold-based catalysts for the
removal of NO in the critical conditions of the cold start phase,
immediately following the start-up of a vehicle engine. We recently
investigated CeO2-supported gold for the reaction of NO with
CO.118–120 The effect of ceria promotion by Al2O3 and the effect of
the support preparation on gold activity for the NO+CO reaction
were examined. CeO2 and mixed oxides of CeO2 with 10 and 20
wt% of alumina were prepared by co-precipitation (CP)117 and by
mechanochemical activation (MA).120 Then 3 wt% amount of gold
was added by deposition-precipitation. According to XRD analyses,
the fluorite structure of CeO2 was obtained in all cases; however, in
the presence of 20 wt% of alumina, broader diffraction peaks
indicative of smaller CeO2 crystallite particles were observed. The
increase of the FWHM of the ceria Raman line was attributed to the
decrease of the particle size of ceria and to the formation of oxygen
vacancies enhanced by the presence of the alumina.121 TPR profiles
of Au/CeO2 and Au/CeO2–Al2O3 were in accord with an increased
reduction of the ceria in the presence of alumina. The catalysts
were tested after different pre-treatments and using different feed
compositions.119 It was established that the type of pre-treatment,
either H2 reduction at 120 °C or air oxidation at 400 °C, had little
influence on the catalytic activity of both Au/CeO2 and Au/CeO2–
Al2O3 systems with a slight improvement in the case of the pre-
reduced catalysts. Using an equimolar NO and CO mixture, with
both types of catalyst the conversion of the reactants started above
100 °C without reaching full conversion in the 50–600 °C tempera-
ture range considered. In accord with the extremely high activity
for the low-temperature oxidation of CO by CeO2-supported gold
catalysts, adding oxygen to the feed gave full conversion of CO to
CO2 without any NO reduction. Considering that hydrogen is
always present in exhaust gases, being generated either by the water
gas shift reaction or by steam reforming,117 the NO+CO reaction
was also performed in the presence of H2. Conversion of both CO
and NO increased with the amount of added hydrogen. The 100%
selectivity to N2 was obtained with gold catalysts supported on ceria

b1469_Ch-14.indd 869 4/8/2013 12:42:36 PM


b1469 Catalysis by Ceria and Related Materials

870 A.M. Venezia et al.

and mixed ceria–alumina supports at a temperature near 200 °C.


Moreover, the presence of water favored the NO reduction. Overall
the MA-prepared samples exhibited lower activity compared to the
corresponding samples prepared by CP. The lower activity corre-
lated with the lower number of oxygen vacancies as indicated by the
FWHM of the Raman ceria line and also by the H2-TPR pattern,
which showed a milder modification of the ceria structure with a
limited number of oxygen vacancies in the MA samples compared
to the CP samples.120 The effect of rare-earth metals (Y, La, Sm, and
Gd) on CeO2-supported gold catalysts for the same reaction was
also reported.122 It was confirmed that M3+ dopants with different
ionic radii and different oxidation states acted as structural
promoters by increasing the number of ceria oxygen vacancies.122
A correlation between the low-temperature ceria reducibility and
NO conversion was established.
In order to elucidate the role of hydrogen in the NO+CO reac-
tion, all the species adsorbed during the reaction in the absence and
in the presence of hydrogen over CeO2 and CeO2–Al2O3 and the
corresponding gold catalysts were studied by FTIR spectroscopy.123
The formation of NCO species was detected on all samples. Unlike
the other supported noble-metal catalysts124 and gold on TiO2,125
where the NCO bands formed over the noble metal and then spilled
over the support, in the present systems NCO formed directly over
Ce3+. In contrast to the results for Au/TiO2 reported in the litera-
ture,125 no direct evidence for the dissociation of NO on the sup-
ported gold catalysts was found for gold over ceria-based supports.
It was concluded that the presence of metallic gold was not neces-
sary for the generation of isocyanates on the supports. In accord
with previous TPR analyses, the presence of gold contributed to the
modification of ceria leading to an increase in the number of Ce3+
ions and oxygen vacancies. The NCO species were formed by a pro-
cess involving the dissociation of NO on the oxygen vacancies of the
support, followed by a reaction between the N atoms lying on the
surface and the CO molecules. The isocyanate species then reacted
with NO producing N2 and CO2. The mechanism is summarized in
the following steps:

b1469_Ch-14.indd 870 4/8/2013 12:42:36 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Catalysts for Air Pollution Abatement 871

Au + CeO2 → Auδ+ … CeO2-x


CeO2-x + yH2 → CeO2-x-y +y… +yH2O
NO + … → Ns+Os
Ns+CO → NCOs
NCOs + NO → N2 + CO2

where … stands for an oxygen anion vacancy and the subscript s


refers to surface-adsorbed species. The role of hydrogen in enhanc-
ing NO and CO conversion was indirect, in the sense of keeping the
catalyst surface, particularly the CeO2, reduced during the course of
the reaction. According to the FTIR results, it was concluded that
catalytic activity of these catalysts was not associated with the gold but
to the formation of Ce3+ ions and therefore to the oxygen vacancies.

14.6 Conclusion
As mentioned at the beginning of this chapter the structural and
electronic properties of CeO2 make this oxide particularly suitable
as a component of catalysts for environmental reactions involving
oxidation steps. Through specific but not exhaustive examples from
the recent literature, it was shown how the presence of CeO2 as a
main active species, as a support, or as a catalyst promoter, contrib-
utes to the enhancement of catalytic activity. Its use as a single-oxide
catalyst, particularly for the CO and VOC oxidation reactions,
requires advanced synthesis of nanostructured materials. The con-
trol of the morphology and the structure of the final CeO2 oxide,
with a high surface area and lattice defects, is important for its elec-
tronic and adsorption properties. Its successful use as a support of
noble metals requires synthetic approaches enhancing the interface
between CeO2 and the metal, favoring the eventual spillover of oxy-
gen from the support to the metal and vice versa. Likewise, as a base-
metal-oxide support, intimate contact between the two oxides with
possible solid solution formation was the best structural arrange-
ment for the promotion of oxygen mobility.

b1469_Ch-14.indd 871 4/8/2013 12:42:36 PM


b1469 Catalysis by Ceria and Related Materials

872 A.M. Venezia et al.

The catalytic involvement of ceria in the surveyed reactions


depended strongly on the particular substrate to be oxidized. The
differences were due to the different mechanisms of the reac-
tions. Indeed the mechanism of CO oxidation implied that the
adsorbed CO was oxidized to CO2 through lattice oxygen, making
the oxygen mobility of ceria the most important property. The
oxidation of a VOC involved at first the adsorption of a molecule
at the surface of the catalyst and then the breaking of several
C–C, C–H, and C–O bonds. In this case, the surface acidity and
adsorption properties in addition to oxygen mobility were quite
important.
With respect to CH4 oxidation, the most important step was the
activation of the C–H rupture. Despite theoretical calculations pre-
dicting the complete dissociation of methane over CeO2 to yield
CO2, from a practical point of view, ceria on its own was much less
active than other transition-metal oxides. Nevertheless, the presence
of ceria in noble metals or other oxide-supported systems was very
beneficial.
In NO oxidation by CO, CeO2 on its own had moderate activity,
which supported the mechanism involving the abstraction of oxygen
by CO molecules and the adsorption of NO on the ceria oxygen
vacancies. The enhancement of activity through a bi-functional
mechanism in the presence of a noble metal or promotion by other
oxides was required to achieve the catalytic efficiency needed for
practical application.
In conclusion it can be stated that CeO2, under different roles,
is and probably will continue to be, the material of choice for cata-
lytic application in the removal of noxious compounds.

Acknowledgments
Support from the European Community through the COST Action
CM0903 and by a NATO grant, ESP CLG. 984160, is gratefully
acknowledged.

b1469_Ch-14.indd 872 4/8/2013 12:42:36 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Catalysts for Air Pollution Abatement 873

References
1. Trovarelli, A., In Catalysis by Ceria and Related Materials, ed. A. Trovarelli,
Imperial College Press, Singapore, (2002), pp. 15–50.
2. Wells, A.F., Structural Inorganic Chemistry, 3rd edition, Oxford University
Press, Oxford, (1961), p. 457.
3. Perrichon, V., Laachir, A., Bergeret, G., Frety, R., Tournayan, L.,
Touret, O., J. Chem. Soc. Faraday Trans., 90 (1994) 773–781.
4. Tsunekawa, S., Sivamohan, R., Ito, S., Kasuya, A., Fukuda, T.,
Nanostructured Materials, 11 (1999) 141–147.
5. Lu J.-L., Gao, H.-J., Shaikhutdinov S., Freund H.-J., Surf. Sci., 600
(2006) 5004–5010.
6. Farmer, J.A., Campbell, T., Science, 329 (2010) 933–936.
7. Kaspar, J., Fornasiero, P., Hickey, N., Catal. Today, 77 (2003) 419–449.
8. Venezia, A.M., Pantaleo, G., Longo, A., Di Carlo, G., Casaletto, M.P.,
Liotta, L.F., Deganello, G., J. Phys. Chem. B, 109 (2005) 2821–2827.
9. Liotta, L.F., Deganello, G., Di Carlo, G., Pantaleo, G., Venezia, A.M.,
Current Topics in Catalysis, 7 (2008) 77–97.
10. Su, Y., Wang, S., Zhang, T., Wang, S., Zhu, B., Cao, J., Yua, Z., Zhang, S.,
Huang, W., Wu, S., Catal. Lett., 124 (2008) 405–412.
11. Royer, S., Duprez, D., ChemCatChem, 3 (2011) 24–65.
12. Schubert, M.M., Hackenberg, S., van Veen, A.C., Muhler, M., Plzak, V.,
Behm, R.J., J. Catal., 197 (2001) 113–122.
13. Nolan, M., Fearon, J.E., Watson, G.W., Solid State Ionic, 177 (2006)
3069–3074.
14. Loschen, C., Migani, A., Bromley, S.T., Illas, F., Neyman, K.M., Phys.
Chem. Chem. Phys., 10 (2008) 5730–5738.
15. Mars, P., van Krevelen, D.W., Chem. Eng. Sci., 3 (1954) 41–59.
16. Aneggi, E., Llorca, J., Boaro, M., Trovarelli, A., J. Catal., 234 (2005)
88–95.
17. Zhou, K., Wang, X., Sun, X., Peng, Q., Li, Y., J. Catal., 229 (2005)
206–212.
18. Valechha, D., Lokhande, S., Klementova, M., Subrt, J., Rayalu, S.,
Labhsetwar, N., J. Mater. Chem., 21 (2011) 3718–3725.

b1469_Ch-14.indd 873 4/8/2013 12:42:36 PM


b1469 Catalysis by Ceria and Related Materials

874 A.M. Venezia et al.

19. Booreskov, G.K., In Preparation of Catalyst 1: Scientific Bases for the


Preparation of Heterogeneous Catalysts. Study in Surface Science and
Catalysis, Vol. 1, eds. B. Delmon, P.A. Jacobs, G. Poncelet, Elsevier,
Amsterdam, (1975).
20. Pan, C., Zhang, D., Shi, L., J. Solid State Chem., 181 (2008) 1298–1306.
21. Vantomme, A., Yuan, Z.Y., Du, G., Su, B.L., Langmuir, 21 (2005)
1132–1135.
22. Sayle, D.C., Maicaneau, S.A., Watson, G.W., J. Am. Chem. Soc., 124
(2002) 11429–11439.
23. Ho, C., Jimmy, C.Y., Tszyan, K., Angelo, C.M., Sukyin, L., Chem. Mater.,
17 (2005) 4514–4522.
24. Kang, M., Song, M.W., Lee, C.H., Appl. Catal. A, 251 (2003) 143–156.
25. Luo, J.Y., Meng, M., Li, X., Li, X.-G., Zha, Y.-Q., Hu, T.-D., Xie, Y.-N.,
Zhang, J., J. Catal., 254 (2008) 310–324.
26. Spadaro, L., Arena, F., Granados, K.L., Ojeda, M., Fierro, J.L.G.,
Frusteri, F., J. Catal., 234 (2005) 451–462.
27. Skarman, B., Grandjean, D., Benfield, R.E., Hinz, A., Andersson, A.,
Wallenberg, L.R., J. Catal., 211 (2002) 119–133.
28. Zheng, X.-C., Wu, S.-H., Wang, S.-P., Wang, S.-R., Zhang, S.M., Huang,
W.-P., Appl. Catal. A, 283 (2005) 217–223.
29. Zheng, X.-C., Zhang, X., Wang, X., Wang, S., Wu, S., Appl. Catal. A,
295 (2005) 142–149.
30. Heck, R.M., Farrauto, R.J., in Catalytic Air Pollution Control, 2nd edi-
tion, eds. R. Heck, R.J. Farrauto, Wiley, New York, (2002), pp. 69–129.
31. Shekhtman, S.O., Coguet, A., Burch, R., Hardacre, C., Maguire, N.,
J. Catal., 253 (2008) 303–311.
32. Monteiro, R.S., Dieguez, L.C., Schmal, M., Catal. Today, 65 (2001)
77–89.
33. Costa, C.N., Christou, S.Y., Georgiou, G., Efstathiou, M., J. Catal., 219
(2003) 259–272.
34. Christou, S.Y., Efstathiou, M., Topics in Catal., 42–3 (2007) 351–355.
35. Wang, B., Weng, D., Wu, X., Ran, R., Appl. Surf. Sci., 257 (2011)
3878–3883.
36. Bera, P., Patil, K.C., Jayaram, V., Subbanna, C.N., Hedge, M.S.,
J. Catal., 196 (2000) 293–301.

b1469_Ch-14.indd 874 4/8/2013 12:42:36 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Catalysts for Air Pollution Abatement 875

37. Haruta, M., Yamada, N., Kobayashi, T., Iijima, S., J. Catal., 115 (1989)
301–309.
38. Scirè, S., Minicò, S., Crisafulli, C., Satriano, C., Pistone, A., Appl. Catal.
B. Environm., 40 (2003) 43–49.
39. Carrettin, S., Concepcìon P., Corma, A., Lopez Nieto, J.M., Puntes, V.F.,
Angew. Chem. Int. Ed., 43 (2004) 2538–2540.
40. Fu, Q., Saltsburg, H., Flytzani-Stephanopolous, M., Science, 301 (2003)
935–938.
41. Casaletto, M.P., Longo, A., Martorana, A., Prestianni, A., Venezia, A.M.,
Surf. Interf. Anal., 38 (2006) 215–218.
42. Casaletto, M.P., Longo, A., Venezia, A.M., Martorana, A., Prestianni, A.,
Appl. Catal. A, 302 (2006) 309–316.
43. Liotta, L.F., Appl. Catal. B, 100 (2010) 403–412.
44. Li, W.B., Wang, J.X., Gong, H., Catal. Today, 148 (2009) 81–87.
45. Patterson, M.J., Agove, D.E., Cant, N.W., Appl. Catal. B, 26 (2000) 47–57.
46. Heynderickx, P.M., Thybaut, J.W., Poelman, H., Poelman, D.,
Marin, G.B., Appl. Catal. B, 90 (2009) 295–306.
47. Arzamenti, G., O’Shea, V.A.D., Alvarez-Galvan, M.C., Fierro, J.L.G.,
Arias, P.L., Gandia, L.M., J. Catal., 261 (2009) 50–59.
48. Baldi, M., Finocchio, E., Milella, F., Busca, G., Appl. Catal. B, 16 (1998)
43–51.
49. Dai Q., Wang, X., Lu, G., Appl. Catal. B, 81 (2008) 192–202.
50. Gutierrez-Ortiz, J.I., de Rivas, B., Lopez-Fonseca, R., Gonzalez-
Velasco, J.R, Appl. Catal. A, 269 (2004) 147–155.
51. Gutierrez-Ortiz, J.I., de Rivas, B., Lopez-Fonseca, R., Gonzalez-
Velasco, J.R., Appl. Catal. B, 65 (2006) 191–200.
52. Balducci, G., Kaspar, J., Fornasiero, P., Graziani M., J. Phys. Chem. B,
101 (1997) 1750–1753.
53. Taylor, S.H., O’Leary, R., Appl. Catal. B, 25 (2000) 137–149.
54. Delimaris, D., Ioannides, T., Appl. Catal. B, 84 (2008) 303–312.
55. Delimaris, D., Ioannides, T., Appl. Catal. B, 89 (2009) 295–302.
56. Tang, X., Li, Y., Huang, X., Xu, Y., Zhu, H., Wang, J., Shen, W., Appl.
Catal. B, 62 (2006) 265–273.
57. Kaptejin, F., Singoredjo, L., Andreini, A., Appl. Catal. B, 3 (1994)
173–189.

b1469_Ch-14.indd 875 4/8/2013 12:42:36 PM


b1469 Catalysis by Ceria and Related Materials

876 A.M. Venezia et al.

58. Kang, C.H.Y., Kusaba, H., Yahiro, H., Sasaki, K., Teraoka, Y., Solid State
Ionics, 177 (2006) 1799–1802.
59. Rao, G.R., Sahu, H.R., Mishra, B.G., Colloid Surf. A, 220 (2003) 261–269.
60. Xiaoyuan, J., Guanglie, I., Renxian, Z., Jianxin, M., Yu, C., Xiaoming, Z.,
Appl. Surf. Sci., 173 (2001) 208–220.
61. Avgouropoulos, G., Joannides, T., Appl. Catal. A, 244 (2003) 155–167.
62. Hu C., Zhu Q., Jiang Z., Zhang Y., Wang Y., Micropor. Mesopor. Pater. 113
(2008) 427.
63. Avgouropoulos, G., Joannides, T., Appl. Catal. B, 67 (2006) 1–11.
64. Heynderickx, P.M., Thybaut, J.W., Poelman, H., Poelman, D.,
Marin, G.B., J. Catal., 272 (2010) 109–121.
65. Balcaen, V., Poelman, H., Poelman, D., Marin, G.B., J. Catal., 283
(2011) 75–88.
66. Menon, U., Galvita, V.V., Marin, G.B., J. Catal., 283 (2011) 1–9.
67. Liotta, L.F., Ousmane, M., Di Carlo, G., Pantaleo, G., Deganello, G.,
Boreave, A., Giroir-Fendler, A., Catal. Lett., 127 (2009) 270–276.
68. Mitsui, T., Matsui, T., Kikuchi, R., Eguchi, K., Top. Catal., 52 (2009)
464–469.
69. Mitsui, T., Matsui, T., Tsutsui, M., Kikuchi, R., Eguchi, K., Appl. Catal.
B, 81 (2008) 56–63.
70. Ordòňez, S., Bello, L., Sastre, H., Rosal, R., Diez, F.V., Appl. Catal. B,
38 (2002) 139–149.
71. Radic, N., Grbic, B., Terlecki-Baricevic, A., Appl. Catal. B, 50 (2004)
153–159.
72. Duprez, D., in Studies in Surface Science and Catalysis, eds. C. Li, Q. Xin,
Vol. 112, (1997), p. 13.
73. Scirè, S., Minicò, S., Crisafulli, C., Satriano, C., Pistone, A., Appl. Catal.
B, 40 (2003) 43–49.
74. Scirè, S., Riccobene, P.M., Crisafulli, C., Appl. Catal. B, 101 (2010)
109–117.
75. Liu, S.Y., Yang, S.M., Appl. Catal., 334 (2008) 92–99.
76. Ousmane, M., Liotta, L.F., Di Carlo, G., Pantaleo, G., Venezia, A.M.,
Deganello, G., Retailleau, L., Boreave, A., Giroir-Fendler, A., Appl.
Catal. B, 101 (2011) 629–637.
77. Kasture, M.B., Patel, P., Prabhune, A.A., Ramana, C.V., Kulkarni, A.A.,
Prasad, B.L.V., J. Chem. Sci., 120 (2008) 515–520.

b1469_Ch-14.indd 876 4/8/2013 12:42:36 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Catalysts for Air Pollution Abatement 877

78. Bonelli, R., Albonetti, S., Morandi, V., Ortolani, L., Riccobene, P.M.,
Scirè, S., Zacchini, S., Appl. Catal. A, 395 (2011) 10–18.
79. Minico, S., Scirè, S., Crisafulli, C., Maggiore, R., Galvagno, S., Appl.
Catal. B, 28 (2000) 245–251.
80. Gelin, P., Primet, M., Appl. Catal. B, 39 (2002) 1–37.
81. Lambert, J.K., Kazi, M.S., Farrauto, R., J. Appl. Catal. B, 14 (1997)
211–223.
82. Liotta, L.F., Di Carlo, G., Pantaleo, G., Venezia, A.M., Deganello, Appl.
Catal. B, 66 (2006) 217–227.
83. Liotta, L.F., Di Carlo, G., Pantaleo, G., Deganello, G., Appl. Catal. B,
70 (2007) 314–322.
84. Yoshida, H., Nakajima, T., Yazawa, Y., Hattori, T., Appl. Catal. B, 71,
(2007) 70–79.
85. Venezia, A.M., Murania, R., Pantaleo, G., Deganello, G., J. Catal., 251
(2007) 94–102.
86. Venezia, A.M., Di Carlo, G., Liotta, L.F., Pantaleo, G., Kantcheva, M.,
Appl. Catal. B, 106 (2011) 529–539.
87. Venezia, A.M., Di Carlo, G., Pantaleo, G., Liotta, L.F., Melaet, G.,
Kruse, N., Appl. Catal. B, 88 (2009) 430–437.
88. Choudary, T.V., Banerjee, S., Choudhary, V.R., Appl. Catal. A, 234
(2002) 1–23.
89. Knapp, D., Ziegler, T., J. Phys. Chem. C, 112 (2008) 17311–17318.
90. Colussi, S., De Leitenburg, C., Dolcetti, G., Trovarelli, A., J. Alloys Comp.,
374 (2004) 387–392.
91. Ramirez-Lopez, R., Martinez, I.E., Balderas-Tapia, L., Catal. Today, 150
(2010) 358–362.
92. O’Connell, M., Morris, M.A., Catal. Today, 59 (2000) 387–393.
93. Santhosh Kumar, M., Eyssler, A., Hug, P., van Vegten, N., Baiker, A.,
Weidenkaff, A., Ferri, D., Appl. Catal. B., 94 (2010) 77–84.
94. Mayernick, A.D., Janik, M.J., J. Catal., 278 (2011) 16–25.
95. Liotta, L.F., Di Carlo, G., Pantaleo, G., Venezia, A.M., Deganello, G.,
Top. Catal., 52 (2009) 1989–1994.
96. Liotta, L.F., Di Carlo, G., Pantaleo, G., Venezia, A.M., Deganello, G.,
Merlone Borla, E., Pidria, M., Appl. Catal. B, 75 (2007) 182–188.
97. Colussi, S., Trovarelli, A., Vesselli, E., Baraldi, A., Comelli, G., Groppi, G.,
Llorca, J., App. Catal. A, 390 (2010) 1–10.

b1469_Ch-14.indd 877 4/8/2013 12:42:37 PM


b1469 Catalysis by Ceria and Related Materials

878 A.M. Venezia et al.

98. Simplício, L.M.T., Brandão, S.T., Domingos, D., Bozon-Verduraz, F.,


Sales, E.A., Appl. Catal. A, 360 (2009) 2–7.
99. Monteiro, R.S., Diegues, L.C., Schmal, M., Catal. Today, 65 (2001)
77–89.
100. Kylhammar, L., Carlsson, P.-A., Skoglundh, M., J. Catal., 284 (2011)
50–59.
101. Becker, E., Carlsson, P.-A., Skoglundh, M., Top. Catal., 52 (2009)
1957–1961.
102. Kylhammar, L., Carlsson, P.-A., Ingelsten, H.H., Gronbeck, H.,
Skoglundh, M., Appl. Catal. B, 84 (2008) 268–276.
103. Grisel, R.J.H., Nieuwenhuys, B.E., Catal. Today, 64 (2001) 69–81.
104. Waters, R.D., Weimer, J.J., Smith, J.E., Catal. Lett., 30 (1995) 181–188.
105. Liotta, L.F., Di Carlo, G., Pantaleo, G., Deganello, G., Catal. Comm., 6
(2005) 329–336.
106. Liotta, L.F., Di Carlo, G., Longo, A., Pantaleo, G., Venezia, A.M., Catal.
Today, 139 (2008) 174–179.
107. Bera, P., Patil, K.C., Jayaram, V., Subbanna, G.N., Hegde, M.S., J Catal.,
196 (2000) 293–301.
108. Roy, S., Hegde, M.S. Madras, G., Appol. Energy, 86 (2009) 2283–2297.
109. Granger, P., Dujardin, C., Paul, J.-F., Leclercq, G., J. Mol. Catal. A, 228
(2005) 241–253.
110. Chafik, T., Kondarides, D.J., Verykios, X.E., J. Catal., 190 (2000)
446–459.
111. Granger, P., Delannoy, L., Lecompte, J.J., Dathy, C., Praliaud, H.,
Leclerq, L., Leclerq, G., J. Catal., 207 (2002) 202–212.
112. Granger, P., Lamonier, J.F., Sergent, N., Aboukais, A., Leclerq, L.,
Leclerq, G., Topics in Catalysis, 16/17 (2001) 89–94.
113. Chen, J., Zhan, Y., Zhu, J., Chen, C., Lin, X., Zheng, Q., Appl. Catal. A,
377 (2010) 121–127.
114. Chen, J., Zhu, J., Zhan, Y., Lin, X., Cai, G., Wei, K., Zheng, Q., Appl.
Catal. A, 363 (2009) 208–215.
115. Wang, Y., Zhu, A., Zhang, Y., Au, C.T., Yang, X., Shi, C., Appl. Catal. B,
81 (2008) 141–149.
116. Cheng, X., Zhu, A., Zhang, Y., Wang, Y., Au, C.T., Shi, C., Appl. Catal.
B, 90 (2009) 395–404.
117. Ueda, A., Haruta, M., Gold Bull., 32 (1999) 3–11.

b1469_Ch-14.indd 878 4/8/2013 12:42:37 PM


b1469 Catalysis by Ceria and Related Materials

Ceria-Based Catalysts for Air Pollution Abatement 879

118. Ilieva, L., Pantaleo, G., Ivanov, I., Venezia, A.M., Andreeva, D., Appl.
Catal. B, 65 (2006) 101–109.
119. Ilieva, L., Pantaleo, G., Mintcheva, N., Ivanov, I., Venezia, A.M.,
Andreeva, D., J. Nanosci. Nanotechnol., 8 (2008) 867–873.
120. Ilieva, L., Pantaleo, G., Sobczak, J.W., Ivanov, I., Venezia, A.M.,
Andreeva, D., Appl. Catal. B, 76 (2007) 107–114.
121. Andreeva, D., Nedyalkova, R., Ilieva, L., Abrashev, M.V., Appl. Catal. B,
52 (2004) 157.
122. Ilieva, L., Pantaleo, G., Ivanov, I., Nedyalkova, R., Venezia, A.M.,
Andreeva, D., Catal. Today, 139 (2008) 168–173.
123. Kantcheva, M., Samarskaya, O., Ilieva, L., Pantaleo, G., Venezia, A.M.,
Andreeva, D., Appl. Catal. B, 88 (2009) 113–126.
124. Bánsági, T., Zakar, T.S., Solymosi, F., Appl. Catal. B, 66 (2006) 147–150.
125. Solymosi, F., Bánsági, T., Zakar, S., Catal. Lett., 87 (2003) 7–10.

b1469_Ch-14.indd 879 4/8/2013 12:42:37 PM


b1469 Catalysis by Ceria and Related Materials

This page intentionally left blank

b1469_Ch-14.indd 880 4/8/2013 12:42:37 PM


b1469 Catalysis by Ceria and Related Materials

INDEX

16
O/18O isotopic exchange 158 CO oxidation 827
deposition-precipitation (DP)
adsorption 506
hydrogen 753 hydrogenation of
N2O 228 crotonaldehyde 798
Ag/CeO2 437, 442, 596, 846 oxidation of methane 859
air/fuel ratio 141 sensors 335
Al2O3/CeO2 789 water gas shift reaction
aldol condensation 802 (WGSR) 325, 327
ammonia 706 XPS 829
anode polarization Au/CeO2–Al2O3 502, 511
resistance 687 autothermal reforming 705
anode reactions 728
anode shielding effect 684 Ba,K/CeO2 609
atomistic computer simulation ball milling 428
247, 318 Bi2O3 3, 587
ceria nanoparticle 273 bioethanol 784
ceria nanotube 261
crystallisation 269, 270 CaO 176
dislocation 256 carboxylation of glycerol 789
energy minimisation 249 Carnot cycle 676
grain boundary 257 catalytic combustion of VOC
oxygen transport 250 432
oxygen vacancies 250, 253 catalytic wet-air oxidation of
thin films 267 organics (CWAO) 48
Au/CeO2 325, 329, 331, 436, Ce0.5Zr0.5O2 20, 32, 53, 58, 59,
439, 451, 453, 466, 470, 477, 178, 417, 435, 439, 451, 453
498, 506, 510, 540, 551, Ce0.6La0.4O1.8 736
787, 843 Ce0.6Y0.4O1.8 736

881

b1469_Index.indd 881 4/8/2013 12:43:58 PM


b1469 Catalysis by Ceria and Related Materials

882 Index

Ce0.7Zr0.3O2–x 644 CO chemisorption 108


Ce0.8Hf0.2O2 402, 453 condensation reaction 803
Ce0.8Pr0.2O2 453 CO oxidation 156
Ce0.8Pr0.2O2–x 666 Cu/CeO2–ZrO2 328
Ce0.8Pr0.20O2–x 722 dehydration of glycerol 792
Ce0.8Sm0.2O1.9 690 electronic conductivity 644
Ce0.8Zr0.2O2 19 H2 chemisorption 62, 71
Ce0.9Gd0.1O1.95 412 high-angle annular dark-field
Ce0.9Gd0.1O1.95–x 669, 685 scanning-transmission elec-
Ce0.9Pr0.1O2 417 tron microscopy (HAADF-
Ce0.65Zr0.35O2 18 STEM) 57, 58, 64–66
Ce0.68Zr0.32O2 196 high-resolution transmission
Ce0.76Zr0.24O2 606 electron microscopy
Ce1–xZrxO2 329 (HRTEM) 56, 58, 67, 313,
Ce2O3 635, 744 330
Ce2Zr2O7 56, 199 high temperature aging 192
Ce7O12 3 κ-like phase 199
CeAlO3 187 mesoporous
Ce-modified ZSM-5 785 ceria-zirconia 312
CenO2n−2m 636 NOx adsorption 236
Ce(NO3)3·6H2O 816 oxidation of VOC 833
CeO1.5 630 phase segragation 70
CeO1.92 636 phosphated Ce–Zr–O 174
CeO1.702 636 pyrochlore-like structure
CeO2 279 56–60, 65, 66, 68, 70
high-resolution transmission synthesis of cyclic
electron microscopy carbonates 788
(HRTEM) 320 temperature-programmed
nanostructures 296, 297, reduction (TPR) 52, 60, 61
308, 322 CeO2–ZrO2/Al2O3 425, 435, 439
NOx adsorption 234 Ce–O bond 26
strain effect 322 CeOCl 204
CeO2–Al2O3 534 CePO4 170
CeO2–HfO2 442 CePt5 alloy 796
CeO2–ZrO2 329, 403, 407, 413, ceria 397, 420, 429, 499
420, 434, 440, 446, 451, 537 0D ceria 297
chemical poisoning by P,Ca 1D ceria 302, 319
and Zn 177 2D ceria 306, 311

b1469_Index.indd 882 4/8/2013 12:43:59 PM


b1469 Catalysis by Ceria and Related Materials

Index 883

3D ceria 311, 316 mesoporous ceria 281, 312


(100) face 500, 507 microemulsion 503
(100) surface 253, 278, 282, nanocrystal 543
752 nanotubes 261
(110) face 477, 500, 507 oxygen diffusion
(110) surface 253, 299 coefficient 280
(111) face 470, 476, 500, oxygen pressurer dependence
507, 516, 545 on x in CeOx 634
(111) surface 253, 265, 275, phase diagram 2, 637
277, 282, 299, 752, 816 photonic crystals 343
(200) surface 299 preferential oxidation
(310) surface 252, 277 (PROX) 327
biocompatibility 336 reaction with P2O5 170
carbon nanotubes 309 reduction of ceria 750
ceria nanocrystal 297 self-assembly 281, 311, 314,
ceria nanorods 275 316, 317
ceria nanotubes 304 sensors 332
conductivity of CeOy 654 sol-gel 504
CO oxidation 277, 327 solubility of hydrogen 647
co-precipitation 400, 503, structural disorder 30
509 superstructures 635
density functional theory synthesis of nanoparticles
(DFT) 287 297, 399
electrolytes 678 temperature dependence
electronic conductivity 653 on x in CeOx 636
electronic structures 24 temperature-programmed
grain boundary 258 reduction (TPR) 325, 647
heat capacity 17 thermal expansion
high-resolution transmission coefficient 627
electron microscopy thin film 254, 259, 266, 279,
(HRTEM) 298, 301, 308, 335, 343
345, 421 ceria nanoparticles 743
high-temperature aging 186 ceria–yttria
hydrothermal synthesis 501 Ce0.93Y0.07O1.96 33
ionic conductivity 652 ceria–zirconia
lattice parameter 627 Calculation of PHase Diagram
lattice stability 17 (CALPHAD) 17
lattice strain in reduced crystal structure 4, 7
ceria 630 DFT calculations 14

b1469_Index.indd 883 4/8/2013 12:43:59 PM


b1469 Catalysis by Ceria and Related Materials

884 Index

electron-density CuO/CeO2 822, 838


distributions 32 cyclic carbonate 788
electronic structures 24
G-x diagrams 22 Deacon reaction 834
neutron powder defect chemistry 143, 626
diffraction 12 defect interaction 651
oxide-ion diffusivity 31 dehydration of glycerol 791
oxygen displacement 14, 27 dehydrogenation reaction 784
phase diagram 16 deposition-precipitation (DP)
Raman bands 13 827
Raman spectrum 24 DFT calculations 752
structural disorder 30 diesel particulate filter
unit-cell parameters 8, 11 (DPF) 568
X-ray powder diffraction 12 diesel particulate matter 566
ceria–zirconia–yttria (CZY) 405 diesel soot oxidation 574
cerium aluminate 187 dimethyl carbonate 786
Co3O4 718 dip coating 701
Co3O4–CeO2 820, 842 doped ceria 254, 325, 344, 402,
CO chemisorption 79, 80 403, 411, 415, 418, 423, 425,
CO disproportionation 151 427, 430, 431, 437, 441, 446,
combustion of methane 850 449, 452, 453, 503, 505, 513,
combustion synthesis 422, 502, 516, 520, 529, 533, 535, 552
865 alkali doping 591
composite membranes 713 Ce0.9Pr0.1O2 229
condensation of acetic acid 800 Ce0.8R0.2O1.9 35
CoOx/CeO2 819 Ce0.9Sn0.1O2 315
CO oxidation 48, 80, 444 Ce0.93Y0.07O1.96 33
core-shell materials 362 CeO2–Pr6O11–Bi2O3 587
Ag@CeO2 367, 372, 385, 388 CeO2–ZrO2–La2O3 194
Au@CeO2 366, 385 ceria-based membrane 708
Ir@CeO2 366, 384, 385 ceria–hafnia 585
Pd@CeO2 367, 371, 373, 388 ceria–lanthana 585
preparation 364, 369, 373 dehydration of glycerol 793
Pt@CeO2 369, 371, 385, 388 effect of cobalt on
Cu/CeO2 328, 331, 404, 425, conductivity 723
436, 439, 442, 450, 466, 470, electronic conductivity 664
477, 550, 745, 866 ionic conductivity 658
water gas shift reaction 325 lattice parameter 628

b1469_Index.indd 884 4/8/2013 12:43:59 PM


b1469 Catalysis by Ceria and Related Materials

Index 885

membrane transmission electron


performances 714 microscopy
Mn doping 594 (HAADF-STEM) 104
MnOx–CeO2 608 high-resolution transmission
reduction 644 electron microscopy
solubility limits 629 (HRTEM) 104
transport properties 722 strong metal/support
interaction 115
electrochemical synthesis 426
electrode kinetics 683 H2 chemisorption 79, 80, 82
electrode reaction 729 H2S 744
electronic conductivity 653 hydrocarbon combustion 48
Eley–Rideal mechanism 151, 827 hydrocarbon oxidation 453
hydrogenation of
fluorite 1 cinnamaldehyde 799
fuel cell efficiency 686 hydrogenation of
crotonaldehyde 795
Gd-doped ceria hydrogen oxidation 737
electrical conductivity 673 hydrothermal 404
electrolyte 684
electronic mobility 666 impregnation 424
membranes 711 inverse catalyst 466, 474
Ni supported anode 700 ionic conductivity 652
reduction and oxidation 674
SOFC anode 736 K/CeO2 594
ultrathin films 674 ketonization of carboxylic
glycerol 789 acids 800
gold supported on ceria based Kröger–Vink 626
materials 79
CO chemisorption 86, 96, La0.5Ce0.5O1.75 503
98, 99, 103, 107 La2Zr2O7 701
Fourier transform infrared lambda sensor 141
81, 95, 97, 98, 100, 111, 114 levulinic acid 800
H2 chemisorption 82, 84, 89,
92 Mars–van Krevelen
H2/D2 exchange 87 mechanism 815, 831
high-angle annular M/CeO2/TiO2 485
dark-field scanning- membrane 710

b1469_Index.indd 885 4/8/2013 12:43:59 PM


b1469 Catalysis by Ceria and Related Materials

886 Index

mesoporous silica 801 oxidation of VOC 835


metal organic deposition oxidative dehydrogenation 784
(MOD) 701 oxide-ion conductivity 37
methane combustion 436, 438 oxide ion mobility 652
methanol 706 oxyfuel processes 705
methyl isobutyl ketone 803 oxygen 701
MgO 786 activity 684
MgO–CeO2 402, 787 diffusion 158
Michael condensation 802 diffusion coefficients 163
microemulsion 414 flux 702
Mn0.86Ce0.14O2 438 nonstoichiometry 640
Mn3O4 836 partial molar enthalpy 640
MnOx–CeO2 835 partial molar free energy 636
permeation membranes 703
Na/CeO2 791 peroxide 581, 604
Nb2O5/CeO2 789 peroxide species 161
Ni/CeO2 454 pressurer dependence on x in
NiO 867 CeOx 634
Ni–SDC 740 spillover 160
nitrogen oxides 223 superoxide 581
N2O 226 superoxide species 161
NO2 225 surface oxygen mobility 166
Ni–YSZ 692, 732 oxygen storage capacity complete
Ni–YSZ cermets 679 (OSCC) 144
noble metals 324, 465, 475, 478, oxygen storage capacity
489, 784, 826 (OSC) 51, 60, 62, 63, 140,
NO+CO reaction 866 323, 398
NO oxidation 603 Ce0.69Zr0.11Bi0.20O1.9 587
NOx abatement 48 CO-OSC 73, 75, 77
CO-pulse injection 150, 155
Onsager coefficients 667 dynamic 53
open-circuit voltage 738 dynamic OSC 152
oxidation of arabinose 805 effect of Fe 183, 184
oxidation of carbon effect of Pb 180
monoxide 752, 814 effect of reducing agents 150
oxidation of ethanol 839 experimental set-up 146
oxidation of propane 841 H2-OSC 75, 77
oxidation of toluene 841 H2 pulse-injection 152

b1469_Index.indd 886 4/8/2013 12:43:59 PM


b1469 Catalysis by Ceria and Related Materials

Index 887

millisecond scale 156 reforming reactions 48


Pd/CexZr1-xO2-d 146 Rh/Ce0.63Zr0.37O2 193
pulse-injection method 143 Rh/CeO2 226, 476, 784
reduction/oxidation N2O chemisorption 228
cycles 53, 54, 56–58, 60, N2O decomposition 226
62, 64, 66, 69, 70, 71 rice-ball nanostructure 597
relation with soot Rietveld analysis 29
oxidation 583 Ru/CeO2 842
static 53
screen printing 701
partial oxidation 48 slurry spraying 701
Pd/Ce0.5Zr0.5O2 182 SO2 deactivation 860
Pd/CeO2 171, 784, 826, 854 sol-gel 409
Pd/Co3O4 855 solid oxide fuel cells
PdO 854 (SOFCs) 389, 675
PdO/CeO2/Al2O3 856 modeling 684
PdO decomposition 857 Pd@CeO2 389
Pd–PdO transformation 856 solution combustion
Pd–Rh TWC 153 synthesis (SCS) 577
perovskites 608 solvated metal atom
physical vapor deposition dispersion 827
(PVD) 701 solvothermal 407
PM10 566 sonochemical 420
polaron mobility 653 soot combustion 385, 438
Pr6O11 855 soot removal 48
preferential oxidation space-charge potential 657
(PROX) 48, 80, 366, 384, 497, spin coating 701
545, 549 spinel 718
Au@CeO2 384 spray pyrolysis 430
Ir@CeO2 384 SrCeO3 719
Pt@CeO2 384 SrTiO3 746
propylene carbonate 787 SrZrO3 701
propylene epoxide 787 steam reforming 739
Pt/CeO2 327, 331, 333, 469, 478, strong metal support
784, 795, 859 interaction (SMSI) 81, 114,
sensors 333 195, 324, 795, 797
Pt–Rh/Al2O3–CeO2 863 sulfur poisoning in
pulsed laser deposition electrodes 732
(PLD) 701 supercritical conditions 407

b1469_Index.indd 887 4/8/2013 12:43:59 PM


b1469 Catalysis by Ceria and Related Materials

888 Index

synthesis gas 706 volatile organic compounds


(VOCs) oxidation 48
Tb4O7 855
temperature-programmed water gas shift reaction (WGSR)
oxidation (TPO) 569, 604, 856 48, 80, 325, 363, 376, 465, 497,
temperature-programmed 531, 745
reduction (TPR) 51, 60, 196, Pd@CeO2 378
229, 524, 529, 820, 836 Pd@CeO2/Al2O3 381
three-way catalyst (TWC) 48, 329, Pt-Au@CeO2 378
814, 861 Pt@CeO2 377
Pd-only 48
transesterification reaction 794 yttria-stabilized zirconia
TWC deactivation 169 (YSZ) 676

Vegard’s law 595, 627 ZrO2 6


volatile organic compounds crystal structure 7
(VOCs) 830 Zr–O bonds 26

b1469_Index.indd 888 4/8/2013 12:43:59 PM

You might also like